Sunteți pe pagina 1din 138

APPLICABILITY OF HYDRAULIC BOREHOLE MINING (‘HBHM’) TO

DIAMONDIFEROUS DEPOSITS

By
Daniel Morgan Boone Beck

Submitted in Partial Fulfillment


of the Requirements for the

Master of Mineral Engineering

New Mexico Institute of Mining and Technology


Department of Mineral Engineering

Socorro, New Mexico


(May, 2016)
ABSTRACT

Diamond exploration is an expensive, time consuming process with only a small


probability of success. Less than one percent of kimberlite pipes contain economic
quantities of diamonds. Due to the lack of exploration success the past decade, diamond
demand is projected to outstrip supply by 2018. There exist a number of previously
discovered diamondiferous kimberlite pipes that are unexploited because of their small
size, low grade, remote location or high strip ratio. Conceptually, Hydraulic Borehole
Mining (‘HBHM’) offers a low cost alternative to conventional mining of diamondiferous
kimberlite pipes. It is ideally suited to exploit these typically modest, vertically oriented
ore bodies that are often not economic to exploit using conventional mining methods. The
technology promises significantly lower capital and operating costs than conventional
mining. Overburden does not have to be removed, underground infrastructure is not
required and massive fleets of earth moving equipment are not required. Production is
immediate, highly selective and in the form of a slurry which further reduces transport and
processing costs. Groundwater sources and flows are not disturbed. Environmental
impacts are greatly diminished. While still a prototype technology, this thesis examines in
detail the potential application of HBHM to the mining of kimberlite pipes. Technical
considerations are reviewed and a conceptual economic model of an ‘idealized’ kimberlite
pipe is developed. Results highlight that the technology has the potential to mine these
high value ore bodies at a fraction of the capital and operating costs of conventional mining
methods and achieve economic rates of return for small, low grade deposits. An action
plan to exploit such ‘stranded’ pipes and unlock their intrinsic value is proposed.
Keywords:

Hydraulic Borehole Mining


Hydraulic Mining
Waterjet Mining
Borehole Mining
Kimberlite Deposits
Kimberlite Pipes
ACKNOWLEDGEMENTS

I wish to acknowledge the support and guidance of the Mineral Engineering Faculty,
namely Professors Chavez, Mojtabai and Razavi. Without their patience, input and
support this thesis would not have been possible. I thank each of them for their
contribution. A special note of thanks goes to Dr. Mojtabai who, as my faculty advisor,
provided invaluable guidance and support.

iii
TABLE OF CONTENTS
Page
LIST OF TABLES …………………………………………………………………...vii
LIST OF FIGURES ………………………………………………………………... viii
LIST (PARTIAL) OF ABBREVIATIONS AND ……………………………..……xi
UNIT CONVERSION TABLE ……………………………………………………..xvi
Chapter 1: Executive Summary ……………………………………………………..17
1.1 Hydraulic Borehole Mining ………………………………………….17
1.2 Diamondiferous Deposits …………………………………………….18
1.3 Diamond Supply ……………………………………………………...19
1.4 Global Demand ……………………………………………………….20
1.5 Hydraulic Borehole Mining For Diamondiferous Deposits ………….20
1.6 Conceptual Economics ……………………….………………………22
Chapter 2: The Diamond Industry …………………………………………………...24
2.1 Overview ……………………………………………………………...24
2.2 Geology …………………………………………………………….....24
2.3 Exploration and Development ………………………………………...27
2.4 Production and Processing ……………………………………………28
2.5 Industry Structure and Value Chain …………………………………..29
2.6 Supply and Demand Outlook …………………………………………30
Chapter 3: Hydraulic Borehole Mining Technology …………………………………31
3.1 History ………………………………………………………………...31
3.2 Methodology …………………………………………………………..35
3.3 Water Jet Considerations ……………………………………………...38
3.4 Water Jet Physics ‘101’ ……………………………………………….39
3.5 Cutting Distances ……………………………………………………...43
3.6 Fluid Mechanics and Nozzle Considerations ………………………….46
3.7 Importance of Pressures and Flow Rates ……………………………...60
3.8 Rock Penetration ………………………………………………………61
3.9 Rock Properties – Physical vs. Structural ……………………………..68
3.10 Pulsating Jets …………………………………………………………..75
3.11 Cavitating Jets ………………………………………………………….79
3.12 Extending the Cutting Range – Introduction of Abrasives ……………83
3.13 Airlift ……………………………………………….………………….91
3.14 System Losses in Delivering the Energy ……………………….…….101
3.15 Mine Unit Stability …………………………………………………...111
3.16 Infrastructure and Equipment ………………………………………...112
3.17 Feasibility Study & Pilot Testing Requirements ……………………..115
Chapter 4: Borehole Mining For Diamondiferous Deposits ………………………….117
4.1 HBHM Applied to Diamondiferous Deposits ………………………..117

iv
4.2 Geological & Mining Parameters for Success ………………………...122
4.3 Targeted Diamondiferous Deposits …………………………………...123
Chapter 5: Conceptual Economics ……………………………………………………126
5.1 Conceptual Economic Model ………………………………………….126
5.2 Anticipated Returns …………………………………………………...127
5.3 Key Takeaways From The Financial Analysis ………………………..128
Chapter 6: Conclusions, Action Plan and Challenges ………………………………..129
6.1 Conclusions ……………………………………………………………129
6.2 Challenges ……………………………………………………………..129
6.3 Action Plan …………………………………………………………….130
6.4 Areas of Future Research ………………………………………………130
REFERENCES ………………………………………………………………………..133

v
LIST OF TABLES
Table Page
Table 1 – Historic Hydraulic Borehole Mining Usage ………………………………. 46
Table 2 – Hydraulic Borehole Mining Performance Benchmarks …………………… 47
Table 3 – Relative Jet Performance for Varying Nozzle Sizes ………………………. 76
Table 4 – Hammer vs. Stagnation for a Pulsed Jet …………………………………. 116
Table 5 –Diamond Exploration Targets …………………………………………….. 189
Table 6 –Conceptual Financial Model ……………………………………………… 193

vi
LIST OF FIGURES
Figure Page
Figure 1: Hydraulic Borehole Mining (‘HBHM’) Configuration …………….… 18
Figure 2: Kimberlite Pipe Formation …………………………………………… 19
Figure 3: Global Diamond Production ………………………………………….. 21
Figure 4: Global Sales of Polished Diamonds ……………………………….…. 22
Figure 5: Open-Pit Kimberlite Mine ……………………………………….…… 24
Figure 6: Barge Hydraulic Borehole Mining For Prior Mined Kimberlite …...… 25
Figure 7: Proposed HBHM Mining Pattern for Two Adjacent Pipes …………... 25
Figure 8: Kimberlite Formation ………………………………………………… 30
Figure 9: Model of Kimberlite Pipe …………………………………………..... 31
Figure 10: Worldwide Diamondiferous Deposits ………………………………... 32
Figure 11: No. of Economic Kimberlites ………………………………………… 34
Figure 12: Hydraulic Borehole Mining (‘HBHM’) ……………………………… 48
Figure 13: Integrated Hydraulic Borehole Mining System ……………………… 49
Figure 14: Barge Hydraulic Borehole Mining For Prior Mined Kimberlite …….. 50
Figure 15: Nozzle Section of Waterjet …………………………………………... 53
Figure 16: Breakup Pattern of a Water Jet ………………………………………. 60
Figure 17: Yanaida (1974) - Jet Dispersion Pattern ……………………………... 61
Figure 18: Yanaida (1974) - Dimensionless Pressure Drop With Distance ……... 62
Figure 19: Typical Profile of Pressure Loss with Increasing Stand-Off Distance .. 63
Figure 20: Typical Profile of Depth-of-Cut with Increasing Stand-Off Distance .. 63
Figure 21: Jet Pressure (Bar) vs. Nozzle Diameter (mm) ………………..…....... 76
Figure 22: Jet Pressure (Bar) vs. Flow Rate (lpm) ……………………….……... 77
Figure 23: Water Jet Nozzle Design Parameters ………………………….…….. 79
Figure 24: Nikonov Nozzle Design …………………………………………….. 80
Figure 25: Variety of Water Jet Nozzle Designs …………………………….…. 80
Figure 26: Sample Flow Straightener Geometries ………………………….….. 82
Figure 27: Comparison of Bad (‘a’) vs. Good (‘b’) Nozzle Design for Two Jet System
………………………………………………………………… 83
Figure 28: ‘Bad’ Nozzle Design - Pressure Profiles for Two Jet System …...…… 84
Figure 29: ‘Good’ Nozzle Design - Pressure Profiles for Two Jet System ………. 84
Figure 30: Lohn and Brent HBHM Nozzle Design ………………………………. 86
Figure 31: Illustration of Rehbinder Water Droplet Hydraulic Lift Force ……….. 93
Figure 32: Geometry of Shape Variation w Standoff Distance & Traverse Rate .... 96
Figure 33: Typical Profiles: Depth Penetration vs. Traverse Rate for Two Rock
Permeabilities ………………………………………………………… 98
Figure 34: Fluid Droplet Collapsing Against a Wall ……………………...….… 118
Figure 35: Schematic Showing Development of a Micro-Jet …..………………. 120
Figure 36: High Speed Photos Showing Stages of Bubble Collapse ………….... 121

vii
Figure 37: Cavitation Bubble Collapse and Pressure Trace …………………..… 122
Figure 38: Design of a Nozzle with Center Body to Induce Cavitation …...…... 124
Figure 39: Basic Abrasive Water Jet Design ……………………………..….… 126
Figure 40: Effect of Impact Velocity on Crack Length ………………………… 129
Figure 41: Berea Sandstone Removed by Impact of Steel Balls of Varying Size.. 130
Figure 42: Effect of Increase in Jet Pressure on Cut Depth …………………..… 133
Figure 43: Effect of Increase in AFR on Depth of Cut in Mild Steel …………... 134
Figure 44: Effect of Higher AFR on Cutting Depth at 3 Jet Pressures on a Mild Steel
Target………………………………………………………………..... 135
Figure 45: Illustration of Simple Airlift Pump ……………………………….… 138
Figure 46: Total Energy of Fluid in Pipe Flow ……………….………………… 139
Figure 47: Flow Regimes ……………………………………………………….. 146
Figure 48: Example Flow Regime Map (Taitel et. al. 1980) …………………… 147
Figure 49: Concept. Airlift Pump and Axial Pressure Distribution ……………. 149
Figure 50: Water Jet Deterioration Structure with Increasing Distance ……….. 157
Figure 51: Energy Losses at Each Stage of HBHM System …………………… 168
Figure 52: Energy Retention (K Joules) at Each Stage of HBHM System …….. 168
Figure 53: Typical Equipment Configuration for HBHM Diamond Operation … 174
Figure 54: Idealized Kimberlite Pipe Showing Economic Limits of Open Pit and
Extended Reach of Jet Mining ………………………………………. 178
Figure 55: Idealized HBHM Hole in Kimberlite Pipe ………………………….. 179
Figure 56: Idealized HBHM Holes in Kimberlite Pipe with Targeted Sections
Highlighted in Red ………………………………………………….. 181
Figure 57: HBHM Mining Pattern (Merlin Diamonds) ………………………… 184
Figure 58: Cross Section of Kimberlite Pipe with HBHM Sections Highlighted (Merlin
Diamonds) ………………………………………………..… 184
Figure 59: African Primary and Secondary Diamondiferous Deposit ………….. 188

viii
LIST (PARTIAL) OF ABBREVIATIONS AND SYMBOLS

E Hydraulic Horsepower of Water Jet (kW)


P Pressure of Water Jet (Bar)
Q Flow Rate of Water Jet (liters/min)
Q Flow Rate of Water Jet (liters/min)
𝑐𝑐1 Empirically determined parameter indicating the efficiency of the water jet.
A greater 𝑐𝑐1 represents less energy loss when the water jet exits from the
nozzle.
𝑑𝑑 Nozzle diameter (in millimeters)
P Pressure of Water Jet (Bar)
𝑉𝑉0 Exit velocity of the water jet (m/sec)
𝑐𝑐2 Empirically determined parameter
P Pressure of Water Jet (Bar)
𝑅𝑅 Reaction Force (Kg)
𝑐𝑐3 Empirically determined parameter.
P Pressure of Water Jet (Bar)
Q Flow Rate of Water Jet (liters/min)
𝑐𝑐4𝑋𝑋1 Product of empirically determined parameters 𝑐𝑐1 and 𝑐𝑐4 .
P Pressure of Water Jet (Bar)
𝑑𝑑 Nozzle diameter (in millimeters)
𝑐𝑐5 Empirically determined parameter
P Pressure of Water Jet (Bar)
𝑑𝑑 Nozzle diameter (in millimeters)
𝑣𝑣 Traverse speed with which the nozzle moves over the surface (in
millimeters per sec.)
𝑐𝑐4𝑋𝑋1/5 Product of empirically determined parameters 𝑐𝑐1 and
𝑐𝑐4 divided by empirically determined 𝑐𝑐5
P Pressure of Water Jet (Bar)
𝑑𝑑 Nozzle diameter (in millimeters)
𝑣𝑣 Traverse speed with which the nozzle moves over the surface (in
millimeters per sec.)
𝑃𝑃 Total pressure (psi)
𝑃𝑃0 Water jet stagnation pressure (psi)
𝑋𝑋𝑐𝑐 Non-dimensional length of the initial core region of the water jet
X Non-dimensional standoff distance
𝑉𝑉0 Exit velocity of the water jet (m/sec)
𝑐𝑐2 Empirically determined parameter indicating the efficiency of the water jet
P Pressure of Water Jet (Bar)
𝑆𝑆 Depth of Penetration

ix
𝑘𝑘 Empirical constant depending on the mechanical properties of the target
material and of the projectile
𝐴𝐴 Cross –sectional areas of the projectile
𝑚𝑚 Mass of the projectile
𝑣𝑣0 Terminal impact velocity
𝑣𝑣𝑐𝑐 Critical velocity at which penetration first occurs
𝑉𝑉𝑉𝑉𝑉𝑉 Volume of Penetration
𝑘𝑘1 Empirical constant depending on the mechanical properties of the target
material and of the projectile
𝑚𝑚 Mass of the projectile
𝑣𝑣0 Terminal impact velocity
ṽ Velocity of grain disaggregation
𝑥𝑥𝑃𝑃 Modified permeability of the rock
𝑢𝑢 Dynamic viscosity of water
𝑝𝑝 Pressure of water in the pores
𝐹𝐹 Hydraulic force that acts against a grain of rock exposed to a water jet due
to viscous drag.
V Volume of the grain of rock
𝒫𝒫 Porosity of the rock grain
𝑝𝑝 Pressure of water in the pores
𝑝𝑝 Pressure at depth of the slot
𝑝𝑝0 Water jet stagnation pressure
𝛽𝛽 Empirical constant for a specific rock type
ɦ Slot depth
𝐷𝐷 Width of the slot
ɦ Slot depth
𝐷𝐷 Width of the slot
𝛽𝛽 Empirical constant for a specific rock type
𝑥𝑥𝑃𝑃 Modified permeability of the rock
𝑝𝑝0 Water jet stagnation pressure
𝑢𝑢 Dynamic viscosity of water
ḹ Average grain diameter
𝑝𝑝 Pressure of water in the pores
T Time of exposure T= d/v, where d is the jet diameter and v is the traverse
velocity.
ɦ Slot depth
𝐷𝐷 Width of the slot
𝛽𝛽 Empirical constant for a specific rock type
𝑝𝑝0 Water jet stagnation pressure
𝑝𝑝𝑡𝑡ℎ Threshold pressure

x
𝑅𝑅𝑠𝑠 Erosion resistance
ḹ Average grain diameter
𝛿𝛿 Average pore diameter
Q Flow Rate of Water Jet (liters/min)
𝐷𝐷 Pipe diameter (cm)
𝑅𝑅𝑅𝑅𝑅𝑅 Reynolds number given by 1116.5•(Q/D)
𝑐𝑐6 Empirically determined constant
P Pressure of Water Jet (Bar)
𝑑𝑑 Nozzle diameter (in millimeters)
𝐷𝐷 Stand-Off Distance (in centimeters)
ɳ Airlift Pump Efficiency (%)
ρ Fluid Density (kg/𝑚𝑚3 )
g Gravitational Acceleration (m/𝑠𝑠 2 )
𝑄𝑄𝐿𝐿 Discharge Volume Flow Rate (𝑚𝑚3 /s)
L Pipe Length (m)
𝐻𝐻𝑠𝑠 Static Depth of Water (m)
𝑃𝑃𝑎𝑎 Atmospheric Pressure (N/𝑚𝑚2 )
𝑄𝑄𝑎𝑎 Air Volumetric Flow Rate (𝑚𝑚3 /s)
𝑃𝑃𝑖𝑖𝑖𝑖 Air Injection Pressure (N/𝑚𝑚2 )

xi
UNIT CONVERSION TABLE

Conversion factors for the most common units used in this text are:
Multiply By:
Meters to feet 3.282
Meters to inches 39.37
Degrees to radians 0.01745
Meters/sec. to miles/hr 2.237
Kilograms to pounds 2.205
Kilograms (force) to newtons 0.1019
Bar to pounds/sq. inch 14.49
Cubic Meters to cubic feet 35.33
Cubic meters to gallons (US) 2.64
Kg/cubic meter to lb/cubic foot 1.686
Liters/Min. to gallons per min. (gpm) 0.264
Joules to kWh 0.2778 x 10−6
Kwh to horsepower 1.34

xii
Chapter 1
Executive Summary

1.1 Hydraulic Borehole Mining (“HBHM”)


The purpose of this thesis is to evaluate the potential application of Hydraulic Borehole
Mining (‘HBHM’) technology to economically mine diamondiferous deposits not
previously accessible with conventional mining methods. HBHM is a prototype mining
technology that incorporates a water jet cutting tool along with a downhole slurry pumping
system. Minerals are recovered through a single borehole drilled from the surface. Water
jets from the mining tool erode the ore to form a slurry which is then continuously pumped
from the base of the tool to the surface in a form suitable for pipeline transfer to a
conventional mill or processing plant. (See Figure 1 below). Crushing and grinding costs
are minimized since the ore is reduced to grain size by the jet stream. Tailings from the
onsite processing plant operations can be pumped into the mined out caverns to control
subsidence and reduce waste disposal volumes.

Figure 1 – Hydraulic Borehole Mining (‘HBHM’) Configuration


(Source: Wikipedia)

HBHM offers a number of important advantages over conventional open pit and
underground mining methods, and can access mineral deposits that are not presently mined
because of technical, environmental, access or economic considerations. Since mining
takes place through a single borehole, no overburden removal is required or expensive
access to drive openings into a targeted ore body. The surface footprint is minimal and
mining can take place in the presence of groundwater without disturbing regional ground
water flows or sources. Environmental impacts are minimal and production is immediate.

17
HBHM is selective and can extract deposits that are inaccessible, small or erratically
mineralized making it ideally suited for the mining of small, stranded diamondiferous
deposits that are not presently economic to exploit using conventional mining methods.

1.2 Diamondiferous Deposits

Diamond Formation
Diamond crystals form at a depth greater than 150 kilometers beneath the earth’s
surface, deep within the mantle of the earth when carbon is exposed to extreme pressure
and very high temperatures. After their formation, diamonds are carried to the surface by
volcanic activity. As this molten mixture of magma (molten rock), minerals, rock
fragments and diamonds approaches the earth’s surface, it begins to form an underground
structure (pipe) that is shaped like a champagne flute or carrot. These pipes are called
‘kimberlites’ or ‘kimberlite pipes’ (see Figure 2 below).

Figure 2 – Kimberlite Pipe Formation


(Source: www.allaboutgemstones.com)
Kimberlite Pipes
Kimberlite pipes, the primary source of mined diamonds, can extend as deep as one
to two kilometers underground. Lamproite pipes, which are created in a similar fashion
but lack the overlying rock cap, are shallower, up to 0.5 kilometers in depth, and typically
have a broader, martini-glass shape.
Kimberlite and lamproite pipes are known as primary diamond sources. Secondary
diamond sources are deposits that have been removed from the primary source (a
kimberlite or lamproite pipe) by natural erosion and eventually deposited in riverbeds,
along shorelines, in glaciers and on the ocean floor. They are also known as alluvial
deposits. Although alluvial deposits account for only 10-15% of the world’s diamonds,
they generally yield a disproportionate number of higher-quality stones.

18
No New Discoveries
In the past decade, no major new diamond deposits have been discovered although
considerable geological exploration work has been carried out. Exploration for kimberlites
is a costly and time consuming process with a small probability of success. Fewer than
7,000 kimberlites have been identified globally. Only 10% of those contain diamonds at
all (i.e. are diamondiferous) and only around 10% of this subset are economic. Hence, only
1% of kimberlite pipes contain diamonds in economic quantities using conventional mining
methods. Presently there are approximately 30 major kimberlites being mined worldwide.

1.3 Diamond Supply

Diamond Supply Constrained


Global diamond supply is dominated by four major producers: ALROSA (Russia),
De Beers (Canada, Botswana & the Democratic Republic of Congo), Dominion Diamond
Corporation (Canada) and Rio Tinto (Australia, Canada) – which account for nearly 80%
of global rough diamond demand. Global diamond reserves are mainly concentrated in
Russia, with 44% of the total and Africa possessing about 33%. In terms of resources, the
picture is very different with 14% and 54% attributed to Russia and Africa, respectively.
Global production is approximately 128 million carats per annum (FY 2012). Global
reserves and resources are estimated to be approximately 3.2 billion carats.

Production Peaking
With supply constrained due to lack of discoveries, and only 11 new mines
expected to start production by 2023 yielding an incremental 18 million carats of additional
supply (vs. production of 128 million carats in 2012), it is unlikely that production in the
medium term will surpass its pre-financial crisis peak in 2005 of 177 million carats.

200
176
180 168 163
160
128 128
Millions of Carats

140 120 123


120
100
80
60
40
20
0
2006 2007 2008 2009 2010 2011 2012

Figure 3 – Global Diamond Production, Millions Carats (Bain: 2013)

19
1.4 Global Demand

Global Demand Robust


Global demand in 2012 was mainly driven by the key consumers: the U.S. (39%),
China (15%) and India (11%). The global diamond industry grew at a CAGR of 9% in
terms of sales in 2010-2012. Incremental demand growth was driven by China and India,
which saw strong economic growth during this period.

Figure 4 – Global Sales of Polished Diamonds, US$ Billions (Bain: 2013)

Incremental Demand Growth from China and India


The key theme for the coming decade, according to Bain (2011 & 2013), is demand
growth in the emerging markets, particularly China and India. Demand in these two
countries, driven by growing wealth and expansion of the middle class, is expected to
represent 37% of total global diamond consumption by 2023, vs. 26% in 2012. At the same
time, further growth elsewhere is expected in line with global GDP gains. Bain forecasts
that demand for diamonds thru 2023 will grow at a CAGR in excess of 5%, eventually
outpacing supply by 2018. As supplies tighten, price increases will take hold thereby
stimulating increased emphasis on developing incremental supply sources.

1.5 Hydraulic Borehole Mining For Diamondiferous Deposits

Mining of Kimberlite Pipes


Mining of kimberlite pipes typically commences by conventional open-pit
methods. The relatively low grade of diamondiferous kimberlites requires that a large
amount of material be removed relative to the number of diamonds recovered. Once the
surface deposits have been exhausted, shafts or declines are sunk to access the deeper ore.
This is more costly than open pit mining. The industry as a whole is transitioning to

20
underground mining as surface deposits are depleted and the discovery of new deposits
lags. Figure 5 below shows a typical open-pit kimberlite mine.

Figure 5 – Open Pit Kimberlite Mine.


(Source: sites.google.com)

Often a kimberlite pipe must be abandoned once the surface ore has been depleted
as it is uneconomic to access the deeper ore due to the higher capital and operating costs
implicit in underground mining. This is where HBHM could play an important role and
provide access to these stranded deeper zones.

Merlin Diamonds - HBHM Pilot Test – Results Unknown


A pilot test is presently being conducted by Merlin Diamonds Ltd. on the
abandoned Merlin kimberlite cluster in the Northern Territory of Australia. The previous
owner (Rio Tinto) mined the surface ore but abandoned the deeper zones as they were
deemed uneconomic by means of conventional underground mining. Merlin is planning
to extract the previously abandoned deeper zones by way of HBHM technology. Similarly,
there are numerous examples of un- or underexploited pipes because of their remote
location, small reserves, low grade, narrow surface areas, erratic geometries, environmental
considerations, etc. Such unexploited or “stranded” pipes are potential candidates for
economic exploitation by HBHM methods.

Exploiting Deeper Reserves with HBHM Technology


Below is a rendering of HBHM being employed to extract deeper ore from a
kimberlite pipe whose surface ore was previously mined out using conventional open-pit
mining methods (i.e. the Merlin pilot test currently underway). Drilling is from a barge-
mounted floating rig operating at the bottom of the flooded pit.

21
Figure 6 – Barge Mounted Hydraulic Borehole Mining For Prior Mined Kimberlite
(Source: www.jetmining.com)

Figure 7 below shows the proposed HBHM mining pattern for exploitation of
deeper zones for two adjacent and abandoned kimberlites in the Merlin cluster.

Figure 7 – Proposed HBHM Mining Pattern For Two Adjacent Pipes


(Merlin 2012)

HBHM Still a Prototype Technology


HBHM is still a prototype technology and has not yet been employed on a
commercial scale. In the case of kimberlite pipes, the technology appears (on paper) to
provide an alternative low cost mining approach to these modest, vertically oriented ore
bodies whose scale is such that they do not support a more traditional open pit/underground
development approach. There are however a number of unanswered questions that can
only be addressed through more wide spread application of the technology, e.g. maximum
throughput tonnages, continuity of feed, depth limitations, ore type limitations with respect
to densities, shear strength, hardness, etc.

1.6 Conceptual Economics

Advantages of HBHM Technology


Conceptually, HBHM offers a low cost alternative to conventional mining of
kimberlite pipes. It is ideally suited to exploit these typically modest, vertically oriented

22
ore bodies that are often not economic to exploit using conventional mining methods. The
technology promises significantly lower capital and operating costs than conventional
mining. Overburden does not have to be removed, underground infrastructure is not
required and massive fleets of earth moving equipment are not required. Production is
immediate, highly selective and in the form of a slurry which further reduces transport and
processing costs. Groundwater sources and flows are not disturbed. Environmental
impacts are greatly diminished.

Conceptual Economic Model


To better assess the economic potential of the technology, we have developed a
conceptual financial model of HBHM applied to a theorized kimberlite pipe. The model
assumptions are based on the limited industry pilot test data available and numerous
assumptions by the author. While admittedly fraught with potentially erroneous
assumptions, our conceptual financial model shows that HBHM technology has the
potential to be a game changer. Conceptually it allows the profitable mining of
diamondiferous deposits of significantly smaller size and lower grade than currently viable
with existing mining technology.

Key Takeaways – HBHM Is Potential Game Changer


Key takeaways from the conceptual HBHM economic model are as follows:
• Capital costs for an HBHM operation are significantly lower than for a comparable
conventional diamond kimberlite operation. In the conceptual case, initial capital
costs were estimated at US$22 million for a two rig configuration which mines and
processes approximately 500,000 tonnes per annum of ore yielding production of
approximately 141,000 carats per annum.
• Attractive financial rates of return can be achieved for relatively modest scaled
operations (NPV-US$80 million; IRR-79%). Such small scale operations (i.e.
500,000 tonnes per annum of ore throughput) would be sub-economic by
conventional mining methods.
• Economic rates of returns are achievable for low grade deposits. Sensitivity
analysis suggests that an attractive IRR of 17% can be achieved with a grade as low
as 18 CPHT (vs. Base Case of 30 CPHT). This grade would likely be uneconomic
for even large-scale conventional operations.
• The ability of HBHM technology to render economic smaller sized and lower grade
kimberlite deposits represents a unique business opportunity to exploit such
‘stranded’ pipes and unlock their intrinsic value.

23
Chapter 2
The Diamond Industry

2.1 Overview

Valuing Diamonds
Natural diamonds are valued mostly for their unparalleled hardness and durability,
luster (optical dispersion) and thermal and electrical conductivity. There are four key
parameters that determine the value (and hence the price) of the diamond. First, the size
of the diamond remains of paramount importance as large diamonds are a scarce geologic
occurrence. Second, the shape of the diamond plays an important role: the more ‘cuttable’
the diamond, the higher the yield and the higher the final recovered value. Third the clarity
of the diamond plays an important role, as stones with no or very small internal defects
command a higher price. Fourth, the color. Colorless diamonds are valued highest except
for the very rare naturally occurring yellows, pinks, blues, etc.

Gem vs. Industrial Quality


Diamonds can be divided into two groups: gem quality and industrial diamonds.
Despite the distinction, there are no clear criteria to make the separation. It is rather
dependent on market conditions and preferences. Gem quality diamonds are acceptable
for the production of polished diamonds and are highly valued for their size, color, quality
and shape. Industrial diamonds lack the best of these attributes and are instead mainly used
for grinding, drilling and polishing. The majority of diamonds produced are gem-quality
and their share is expected to grow.

2.2 Geology

Diamond Formation
Diamonds form deep within the mantle of the earth when carbon is exposed to
extreme pressure and very high temperatures. Volcanic rock formations such as kimberlite
or lamproite pipes serve as pathways that convey the fragments of rocks and crystals from
the mantle to the surface. (see Figure 8 below).

24
Figure 8 – Kimberlite Formaton
(Source: www.awesomestories.com)

Kimberlite Pipes
After their formation, violent eruptions carry diamonds to the surface of the earth
along with magma, minerals and rock fragments. As the magma approaches the earth’s
surface it begins to form an underground structure (pipe) that is shaped like a champagne-
flute or carrot. These pipes are called ‘kimberlite pipes’ (see Figure 9 below). Kimberlite
pipes often lie underneath shallow lakes formed in inactive volcanic calderas or craters.
They can extend as deep as 1 to 2 kilometers beneath the surface.

Figure 9 – Model of Kimberlite Pipe


(Source: www.secretgeek.net)

25
Lamproite Pipes
Diamond rich lamproite pipes are extremely rare. To date, the only economically
viable diamond-bearing lamproites have been discovered in Western Australia. They are
created in a similar manner to kimberlite pipes, except that boiling water and volatile
compounds contained in the magma act corrosively on the overlying rock, resulting in a
broader cone of eviscerated rock at the surface. This yields a broader, martini-glass shaped
deposit versus the kimberlite’s narrower champagne flute shape.

Distribution of Kimberlite Pipes – Africa is King


The distribution of kimberlite pipes is disproportionately in Africa, Russia, Canada
and Australia. See Figure 10 below. Africa is particularly target rich holding 54% of
known diamond resources placing it significantly ahead of Russia, the second largest
source of diamond resources with 14%.

Figure 10 – Worldwide Diamondiferous Deposits


(Source: www.info-diamond.com)

Alluvial Sources Provide High Quality Stones


Kimberlite and lamproite pipes are known as primary diamond sources. Secondary
diamond sources are deposits that have been removed from the primary source (a
kimberlite or lamproite pipe) by natural erosion and eventually deposited in riverbeds,
along shorelines, in glaciers and on the ocean floor. These are known as alluvial deposits.
Although alluvial diamonds account for only 10-15% of the world’s diamonds, they
represent a disproportionally high percentage of gem quality diamonds. This is because
only the hardest of alluvial diamonds survive the millions of years of erosion and transport
from their primary sources.

26
2.3 Exploration and Development

Diamond Exploration is an Expensive and Risky Business


Diamond exploration is a time consuming, capital intensive and risky process with
a low probability of success. The high risk element of diamond exploration is primarily a
reflection of the geology which hosts the product. Unlike metal deposits which often have
large low-grade ‘aureoles’ to aid discovery, diamonds form in relatively small pipes which
leave their surrounding host rock untouched. This makes diamond exploration akin to
finding the proverbial needle in a haystack.

Only 1% of Known Kimberlites Are Economic


There are approximately 7,000 known kimberlite and lamproite pipes. Only 10% of
these are diamondiferous (i.e. contain diamonds). Of these, only 10% contain diamonds
in economic quantities. Hence, only 1% of known kimberlites/lamproites are economic.
Presently there are approximately 30 major kimberlites being mined worldwide.

Number of Kimberlites
7000
7000
6000
5000
4000
3000
2000 700 70
1000
0
Known Kimberlites - Diamondiferous Economic
7000 Kimberlites - 10% Kimberlites - 1%

Figure 11 – No. of Economic Kimberlites


(Bain: 2013)

Exploration Process – Surface Sampling & Drilling


The first stage in diamond exploration is the identification of an area or region
having the geological potential to host diamond deposits. After a region deemed to be
prospective has been selected, specific target areas must be defined. This is generally
accomplished through indicator mineral sampling, an exploration technique used
throughout the world by companies searching for diamonds. Once targets have been

27
identified and prioritized, a drill rig is positioned above the target and a hole is drilled. As
the drill penetrates to depth, a continuous sample of the intersected rock is cored and
retrieved for sampling. If the drill core is not kimberlite, drilling of the target is
discontinued. If the target is a kimberlite, sampling for diamonds is undertaken. If the
samples identify sufficiently large concentrations of diamonds, a bulk sampling program
is undertaken to estimate the grade and average value of the diamonds and ultimately the
economic value of the kimberlite pipe. Sampling a wide area of the pipe is necessary to
secure a representative understanding of the diamond content of the ore body.

Expensive Feasibility Study Required to Determine Project Viability


If the grade and diamond value for a deposit are encouraging, a feasibility study is
commissioned to determine the economic viability of a mining operation. The capital costs
to develop a diamond mine are considerable and generally in the range of US$1-2 billion.
Completion of a full feasibility study and related technical evaluation will typically require
2-3 years at a cost of US$50-100 million. If the feasibility study is positive, then permits
and financing must be secured prior to the start of construction which typically takes 2+
years. The time from first discovery to first production is often 10-15 years.

2.4 Production & Processing

Mining of Diamonds Is Relatively Straight Forward


The rough diamond production process consists of mining, crushing, ore
processing, manual rough diamond sorting, processing and valuation. First, ore is mined
by open-pit or underground methods. Open-pit mining is carried out by simple excavation
or by blasting ore seams and transporting ore to the beneficiation plant. Underground
mining is employed in deeper zones, typically after the open-pit resources have been mined
out. Underground mining is carried out via simple blasting of ore seams with the ore being
transported via underground equipment to the main shaft and lifted to the surface for further
transportation to the beneficiation plant.

Processing of Diamonds is More Complex


The next significant, and more complicated, stage of rough diamond production is ore
processing. Processing varies from mine-to-mine but typically the mined ore is sent to the
crushing mill where the ore is crushed and reduced in size, typically to 30 mm. Ore is
then scrubbed and screened, where it is separated into three sizes. Oversized material (>15
mm) is further reduced and undersized material (<1.5 mm) is rejected to the tailings dump,
as diamonds in this undersized ore cannot be extracted profitably. Ore between 1.5 mm
and 15 mm is conveyed to a Heavy Media Separation (‘HMS’) Plant Feed Stockpile. The
majority of the material in the HMS stockpile is kimberlite ore. It contains diamonds and
some other high density minerals. The ore is then processing in a cyclonic separation plant.

28
Heavy media consisting of ferrosilicon powder mixed with water is used to separate the
kimberlite ore from the diamonds and heavy minerals which sink to provide a diamond
rich concentrate. In the recovery plant, the diamond concentrate is fed through a series of
X-Ray sorters. Diamonds fluoresce when exposed to x-rays. Sensors detect the flashes of
light emitted by the diamonds. These send signals to the microprocessor that fires an air
blaster valve at the appropriate moment, blowing the diamonds into a collection box. The
diamonds are acid cleaned, washed, weighed and transferred to the marketing department
where they are sorted, valued and prepared for sale.

2.5 Industry Structure & Value Chain

Diamonds: A Long Journey from Earth to a Bride’s Finger


It is a long journey for a diamond from the earth to a bride’s finger. The industry,
and its corresponding value chain, is structured into three distinct stages: upstream, middle
market and downstream. At the upstream stage, exploration, development and mining of
diamonds takes place. It is then sold by the producer to a middle-market participant who,
in turn, cuts, polishes and assembles the diamond into jewelry. The diamond then moves
to the downstream segment, the retailer who, in turn, sells the finished product to the
customer. As the diamonds pass through the three stages, the price per carat is constantly
increasing, multiplying eightfold by the time the diamond is sold to the retail customer.
Much of the value multiplication is not, however, profit but rather waste incurred from the
transformation of a rough diamond into a finished and polished stone. On average, five
carats of rough diamonds is required to produce one carat of polished gem. Nearly half of
the loss is in the form of low value industrial grade diamonds, which must be discarded at
nominal value to industrial buyers, while the balance is lost in the physical process of
cutting and polishing.

Upstream Has the Best Margins


The upstream segment (De Beers, ALROSA, Rio Tinto and Dominion) is the most
attractive part of the value chain with profit margins of 16-20%. This is followed closely
by the downstream retailer segment (e.g. Tiffany, Cartier, Graff, Harry Winston, etc.) with
margins of 11-14%. Last and least profitable is the middle market (cutters, polishers,
jewelry manufacturers) where profit margins are a slender 1-8%. A brief summary of each
of the three industry stages is as set out below.

Upstream – Highest Margins; Dominated by Four Producers


Simply put, in the upstream stage, diamonds are mined from the ground, then sorted
and sold by producers either through long- or short-term contracts or auctions. Long-term
contracts and auctions account for about 65% and 30%, respectively of rough diamond
sales. The world’s supply of diamonds comes from relatively few sources – only 11

29
deposits account for 60+% of world production. Four producers dominate the industry, De
Beers, ALROSA, Rio Tinto and Dominion. Collectively they account for nearly 80% of
the upstream segment’s worldwide production.

Middle-Market – Highly Fragmented, Low Margins


The middle market is where the diamonds are cut, polished and made into jewelry.
This a highly fragmented segment of the value chain populated by tens-of-thousands of
firms, of all sizes, predominately located in low labor cost centers like India and China.
Margins are thin and competition is fierce. Their main challenge is to remain cost
competitive while, at the same time, maintaining access to supplies of rough diamonds
which are increasingly scarce and sought after.

Downstream – Many Players but High Margins


The downstream segment consists of selling diamonds and diamond jewelry to retail
customers. It is a high margin segment of the industry and like the middle market highly
fragmented with numerous players. As a general rule, the diamond jewelry market closely
tracks the overall luxury goods market.

2.6 Supply and Demand Outlook

Diamond Shortages from 2018 Onward


According to Bain’s 2011 and 2013 Diamond Reports, global demand for rough
diamonds is expected to grow in line with historical trends through 2023. This translates
into a robust compound annual growth rate (‘CAGR’) of 5.1% driven largely by the pace
of middle-class expansion in China and India and steady growth in the global economy.
The supply side is less robust given the lack of recent discoveries. Bain’s forecast calls for
rough diamond production to peak at 169 million carats in 2018, below the pre-crisis peak
of 177 million carats in 2005, and then to decline to 153 million carats in 2023. It is likely
that demand will outpace supplies from 2018 onward, implying higher prices at every stage
of the value chain from this date onward. As diamond prices increase, increased energies
will be focused on finding incremental supply sources.

30
Chapter 3
Hydraulic Borehole Mining Technology

3.1 History

Hydraulic Mining Started in California Gold Rush


The genesis of HBHM dates back to the 1800s when a high pressure jet of water
sprayed through a monitor was used to mine surface ore and gravels in the California gold
rush. The technology then migrated to Canada and Australia. The earliest incarnation of
a sub-surface HBHM tool was embodied in a patent issued to Clayton in 1932 which
proposed a water jet to fragment rock adjacent to a borehole and a downhole slurry pump
to lift the ore slurry to the surface. Patents for more advanced and integrated systems were
issued to Ashton in 1950, Quick in 1955, Fly in 1964, Pfefferle in 1965, Wenneborg in
1973, Archibald in 1974 and Brunelle in 1978.

First HBHM Prototypes Tested in 1970’s


The system patented by Fly was built and used until the early 1970s to excavate
cavities in sandstone and shale to a depth of 120 meters. In 1970, the Reynolds Metals
Company funded a three month pilot test of the Fly designed apparatus at an open pit
bauxite mine in Benton, Arkansas. The test was unsuccessful because of inconsistent and
low production rates. The system patented by Wenneborg was implemented by FMC and
tested in phosphate ore in eastern North Carolina. A novel aspect of this tool was that it
provided a method for drilling the borehole as well mining the ore adjacent to the borehole.
Previous systems required a predrilled and cased borehole. FMC tested this system but
opted not to commercialize it. Archibald’s design was built by Marconaflo and pilot tested
on uranium ore in Wyoming and tar sands in California.

U.S. Bureau of Mines Takes the Lead


During the 1970’s and 1980’s, the United States Bureau of Mines (‘USBM’) took
the lead in advancing HBHM technology under direction of Dr. George Savanick. USBM
built an experimental HBHM tool in the 1970s which was tested on: (i) uraniferous
sandstone ores in cooperation with Rocky Mountain Energy in Wyoming; (ii) oil sands in
California in cooperation with Century Oil and (iii) phosphate deposits in Florida in
cooperation with Agrico. As a follow-up to the encouraging pilot tests, Rocky Mountain
Energy built a commercial-scale HBHM tool incorporating some improvements over the
USBM prototype. The tool was designed with a high pressure educator system to flush the
cuttings from the jet back to the surface. Water was pumped at a high flow rate down the
tool and back up the return line, past an opening. At the opening, a vacuum was created
causing the cuttings to be carried up inside the tool to the surface. This design required the
pumping of high pressure water to the jet nozzle and to the educator. The tool was tested

31
at the Bear Creek uranium mine in Converse County, Wyoming. The test was discontinued
due to a sharp drop in uranium prices in the 1980s.

Successful Florida Phosphate Pilot Test


The mining tool developed by Rocky Mountain Energy was used by the Agrico
Mining Company in a pilot scale borehole mining test in St. John’s County, Florida in the
mid 1980’s. USBM provided funding assistance and technical input. The pilot test mined
1,700 short tonnes of phosphate, at a rate of 40 tonnes per hour, at a depth of 100 meters.
The HBHM tool operated in a submerged cavity and produced a cavity radius of five
meters. An air shroud was subsequently introduced increasing the cutting radius to six
meters.

U.S. Bureau of Mines Funding Dries Up


Due to a reduction of USBM’s research budget for HBHM technology, and lack of
industry interest, no further developments in HBHM were made until the 1990s when
Layne Christensen Company (‘Layne’) undertook, on behalf of a number of mining
companies, kimberlite bulk sampling on various diamond deposits in Canada. Layne
constructed a fleet of hydraulic top drive rigs to provide continuous large tonnage samples
with controlled minimum breakage of diamonds. Refinement of airlift technology, and the
capacity to slow velocities in a controlled manner as the kimberlite returns to surface, was
a key feature of the technology development undertaken by Layne.

Layne Christensen Picks Up the Slack


In 1996, Layne built a small HBHM system for mining deep vertical coals at the
Gregg River Coal Mine in Alberta Canada. This project, in cooperation with Cominco
Engineering, used airlift technology to flow coal to the bit during continuous mining
operations. Tested rates of production were as high as 40 tonnes per hour through a 4”
diameter mining system. Following this successful pilot, a further pilot test was
undertaken, in cooperation with Cogema, to mine uraniferous ore at Cluff Lake,
Saskatchewan. Additional pilot testing and refinement of the technology for use in
uranium mining was undertaken with Cameco, Areva and Dennison. Cameco subsequently
developed a HBHM system for use at Cigar Lake, Saskatchewan. The Cigar Lake
adaptation uses a jet miner which lets the cuttings drop to the bottom of the hole to be
mechanically lifted by a tram and elevator. The ore is located at 150-300 meters where
economics and intense radiation levels preclude conventional mining methods.

Parallel Developments in Russia


In the 1980’s the Russian Drilling Federation adapted the USBM technology
utilizing a downhole eduction jet pump and sliding jet monitor. The system’s limitation is
that it does not allow for submerged cutting of ore under water and importantly requires

32
significantly more energy to operate. Under the Russian design, a second independent high
volume and high pressure pump is required to create the eduction energy for the downhole
pump. By contrast, the advanced USBM design utilizes water density reduction to create
lift. HBHM technology has been pilot tested in the former Soviet Union on various
deposits of iron ore, uranium ore, titanium-zirconium ore, deep-seated gold placers and
amber. In 2011, the Russian adaption of the USBM system was used to extract a bulk
sample at the Emily Manganese deposit located in Emily, Minnesota. Testing is still
underway.

HBHM Enters Hibernation Phase – No Further Progress


In North America, there was a long pause in HBHM technology experimentation
until early 2012 when Kinley Exploration LLC (‘Kinley’) completed a technical and
economic evaluation of HBHM for the Hansen Uranium Deposit located northwest of
Canon City, Colorado. Kinley’s leadership team comprised former Layne executives
(including Colin Kinley) involved with the development of HBHM technology for
continuous bulk diamond sampling and George Savanick who directed USBM efforts at
advancing HBHM technology during the 1970s and 1980s. Kinley’s desk-top study, which
did not involve field tests, indicated a single borehole mining unit could produce
approximately 500,000 lbs. of U3O8 per annum at a competitive cost of US$27 per lb
U3O8. Production rates could be readily increased by operating multiple borehole mining
units. No further progress toward implementation was made due to a lack of funding.

Merlin Diamonds Commences Pilot Test – Results Unknown


Also in 2012, Merlin Diamonds Ltd. (‘Merlin’), an Australian listed diamond
explorer, engaged Kinley’s Australian affiliate (Jet Mining Ltd.) to undertake a feasibility
study and pilot test of HBHM on their Merlin cluster of kimberlites. Merlin is the owner
of the Merlin diamond mine located in the Northern Territory of Australia. It comprises
14 kimberlite pipes, grouped in four clusters. Nine of these pipes were subject to open-pit
mining by Rio Tinto over a five year period commencing in 1998. The operations ceased
in 2003 when all the ore accessible by conventional surface mining was exhausted. Kinley
prepared a feasibility study in late 2012 confirming the economic viability of HBHM
technology to selectively mine deeper zones of interest. Trial mining commenced in
August 2013. Early results were promising with an announcement by Merlin on 17
October 2013 that read, in part, as follows:

“Merlin Diamonds Ltd would like to announce an update to its trial borehole
mining operations at the Merlin Diamond Mine in the Northern Territory. The borehole
mining rig has been operating for over one month and has achieved success in a number
of key areas. The hydraulic jetting tool that cuts the kimberlite material at depth has been
proven to effectively cut the weathered kimberlite and is able to produce diamond bearing

33
ore suitable for lifting via the mining rods. The hydraulic lifting system has been proven to
lift material to the surface of the pit and is able deliver ore to the shaker screen located on
the ground surface adjacent to the pit. The processing plant has also achieved nameplate
capacity of 75t per hour.
Production rates experienced from the borehole mining technique require
optimising to guarantee maximum recovery and profitability through mining, and full
utilisation of the processing plant capabilities. The Company is now at a stage to complete
engineering work that will increase this rate of production. Production rates through the
borehole mining technique have been attained at other diamond deposits and Merlin’s
engineering and feasibility team will spend the coming months to complete the necessary
work. Given the above the Company has decided to enter into a production hiatus until this
engineering work is complete to ensure maximum shareholder return. “

No further updates have been provided as of the date of this report. Final results
of the trial mining exercise will be of great use in further assessing the viability of HBHM
technology in general and for its potential applicability to the selective mining of
diamondiferous deposits.

A Historical Snapshot of HBHM Usage


A historical snapshot (incomplete) of current and past HBHM usage is as shown
below in Table 1.

Year Project Location Depth Ore Type


2013 -
Present Merlin Diamonds NT, Australia 100-300M Kimberlite
2010 -
Present Emily Manganese Emily, Minnesota 200 M Manganese

2003 CBM Stimulation West Arkansas 250 M Coal

2003 CBM Stimulation Powder River, WY 300 M Coal

2000 CBM Stimulation Zimbabwe 600-800 M Shale

1996-Present Water Stimulation Columbia 100-150 M Sandstone

1996 Coal Mining Alberta, Canada 400 M Coal

1996 Uranium Saskatchewan, CA 100 M Sandstone

1990's Diamond Sampling Canada 0-400 M Kimberlite

1989-Present Diamonds Archangel, Russia 80-800 M Kimberlite

1992-1994 CBM Stimulation New Mexico 1 KM Shale

1988-Present Iron Ore Kursk, Russia 400-800 M Oxide Ores

1974 -Present Baltic Amber Russia West 15-30 M Quartz Sand

1987-1991 Kolubara Serbia 20-40 M Quartz Sand

1970-1990 Uranium West Kazakhstan 100 M Clay Stone

1980 Phosphate St.Johns, Florida 70-80 M Phophate

1979 Oil Sands Taft, California 25-50 M Oil Sands


Converse,
1976-1977 Uranium Wyoming 25-35 M Uranium

34
1976 Coal Mining Wilkeson, WA 15-30 M Coal
Titanium-
1976-Present Zirconium Ukraine 40-100 M Quartz Sand

1978-1982 Gold Bajkal Lake Region 60-80 M Gravel

Table 1 - Historic Hydraulic Borehole Mining Usage

HBHM Historic Performance Benchmark Milestones


A summary of performance benchmarks achieved in a number of U.S. based
HBHM pilot test mining operations is summarized in Table 2 below (Summers, 1995).

Avg. Jet Total


Date Ore Type Depth Pressure Avg. Mining Rate Tons Location
(Meters) (Bar) (TPH)
1976 Coal 5 32 9 48 Wilkerson, WA
1977 Uranium 26 18 6 940 Casper, WY
1979 Oil Sand 39 4 14 990 Taft, CA
1980 Phosphate 57 2 55 830 St. Augustine, FL
1980 Coal 3 32 20 ? Black Diamond, WA

Table 2 - Hydraulic Borehole Mining Performance Benchmarks


(Summers 1995)

Work on HBHM Technology Still Muted


Work toward commercializing HBHM technology has been muted for the past three
decades. With the onset of USBM budget cuts for HBHM research in the 1980s, and
limited follow-on industry interest, progress has been minimal. The only active promoters
of the technology at the moment are Kinley/Jet who are progressing trial mining of the
Merlin kimberlites in Australia and, as part of the Merlin pilot test, experimenting with a
system that allows the drill string to remain stationary while the cutters at the mining head
continuously rotate. Previous USBM inspired systems have all used low torque systems
bolted together which effectively limit milling capacity.

3.2 Methodology

HBHM Has Small Footprint


For any HBHM operation, a relatively small diameter hole is drilled into the target
resource. Once the borehole is drilled through the resource interval, the sidewall is cut and
disaggregated with high pressure water and air. The system then lifts the cut ore to the
surface. HBHM allows the mining of underground ores that cannot be economically mined
by conventional open pit or underground mining methods. Typically, the small footprint
for the drilling rig to drill access to the ore, and the mining rig to cut and lift the ore, requires

35
only short term, temporary access with minimal impact to the environment. Moving
significant overburden is no longer necessary and land use is limited to direct access.

Figure 12 – Hydraulic Borehole Mining (‘HBHM’)


(Source: www.blackrangeminerals.com)

Figure 13 – Integrated Hydraulic Borehole Mining System (‘HBHM’)


(Source: www.blog.kmtwaterjet.com)

36
HBHM Can Operate in Water Saturated Environment
Importantly, HBHM allows operations in submerged or high water flow ore bodies
thereby permitting operators to access deposits previously deemed inaccessible,
uneconomic, or both. Where the surface area is flooded, the drill rig and ancillary support
equipment and materials can be mounted on a barge. See Figure 14 below. With
appropriate drilling procedures and casing, the system permits access through surface or
ground water without disturbing regional ground water flows and sources. Ground water
can be protected from recharge or discharge of inter-aquifer flows and any contamination
from the target when sealed. Mining through a targeted hole, drilled directly and
specifically into the resource, allows economic recovery of targeted deposits. This ability
to selectively mine in a water saturated environment allows high grading without the
dilutive effects of overburden removal.

Figure 14 – Barge Mounted Hydraulic Borehole Mining For Prior Mined Kimberlite
(Source: www.jetmining.com)

HBHM Most Effective in High Strip Ratio Deposits


HBHM is most effective where a specific ore zone can be directly accessed through
a bore hole drilled through uneconomic overburden or in a deposit whose geometry (e.g.
vertical kimberlite pipes with high strip ratio) does not permit efficient mining via
conventional methods. Similarly, the application of HBHM to flat lying ore bodies (e.g.
phosphate, mineral sands, uranium) is advantageous as it avoids the high overburden
removal costs implied by conventional open-pit mining methods. Deposits in high water
table environments also lend themselves to the use of HBHM technology.

37
3.3 Water Jet Considerations

How Water Jets Cut Rock


Water jets disaggregate ore by inducing failure at the weakest point in the target
surface. Much more so than with mechanical tools, it is possible for a water jet to enter
and exploit these weakness planes, thereby removing ore at significantly lower cutting
pressures than otherwise would be required.

Past Studies
The study of cutting rocks using a high velocity jet of water (i.e. water jet) dates
from the 1960’s (Farmer and Attewell, 1964, Brook and Summers, 1968) when the
mechanism and threshold conditions required to break rocks were investigated. Farmer
and Attewell (1964) found that a water jet with insufficient velocity would not be able to
cut rocks but merely wash away the poorly cemented grains from the rock surface. It was
determined that, given the same pressure, a greater nozzle diameter (i.e. greater flow rate)
enables a greater cutting depth (Harris and Mellow, 1974). Further studies by Rehbinder
(1977, 1978, 1979) postulated that water jets, on a micro-scale, create hydraulic lift against
grains on the target surface and wash such grains away so long as the cutting pressure of
the jet exceeds the erosion resistance of the grains. Rehbinder (1979) went further and
developed predictive models of such forces and cutting dynamics. Engin (2012) examined
cutting depths using an abrasive water jet on 42 different natural stone types. His study
found that rock hardness, surface abrasion resistance and density of the rocks were the most
significant rock properties affecting the cutting depth. Water jet impact pressure and
traverse tool velocity were the most significant operating parameters affecting cutting
depth.

Many Complexities – Hard to Model


There are too many variables and complexities in the operation of a water jet system
to allow for a unified theory describing the stream of water from when it is first formed
into a water jet until the time that material is removed from the target surface. Even in the
development of flow from an orifice, the condition of the flow passage upstream and the
shape of the orifice itself will have a significant effect on the jet condition as it leaves the
nozzle and the follow-on energy content and distribution which the jet retains at surface
impact.

Energy Input Costs Must be Measured Against Economic Output


A key to assessing the feasibility of HBHM for a particular ore body is determining
the net hydraulic horsepower and volume required to disaggregate the rock and move it to

38
the return inlet of the mining system. Often the disaggregation of rock is technically
feasible, but the cost of energy input exceeds the economic value of the ore being mined.

Threshold and Cutting Pressure Are Key Variables


A key variable in determining HBHM feasibility is the threshold pressure; i.e. the
minimum pressure required for the water jet to penetrate the target material. This may
also be thought of as the minimum pressure necessary to push the water behind the rock
particles and disaggregate same. The overall HBHM system economics are related to the
practical cutting pressure necessary to disaggregate the ore so the system can return same
to the surface at a fast enough rate and a low enough consumption of energy and cost to
render HBHM economically viable. A useful ‘rule-of-thumb’ is that the cutting pressure
is three-to-four times the threshold pressure required to cut the rock (Rehbinder, 1979).
For typical HBHM applications, cutting pressure is limited to a practical maximum of
5,000 psig given constraints of currently available equipment and components. Ore types
that have a cutting pressure greater than this limit are not suitable candidates for HBHM
technology.

Figure 15 – Nozzle Section of Waterjet


(Source: www.blog.kmtwaterjet.com)

Water Jet Testing of Drill Core Necessary to Determine Cutting Pressure


As part of the feasibility study phase of any HBHM project, drill core samples will
undergo water jet testing to determine threshold and cutting pressures for individual
sections of the target ore body. In the case of kimberlite pipes, the 5,000 psig maximum
limit implies that only softer weathered kimberlite sections will be amenable to selective
mining using HBHM technology. Unweathered sections are likely to be too hard or dense
to be economically cut with a water jet.

3.4 Water Jet Physics ‘101’

Hydraulic Horsepower is a Product of Pressure and Flow Rate

The net hydraulic horsepower of a water jet, E, is the product of the pressure and
the flow rate of fluid at the jet and is a key determinant of the economic viability of the
operation. The hydraulic power of a water jet can be expressed as follows:

39
E = ⦗(1.666•P) • Q⦘/1000 (1.1)

Where:

E Hydraulic Horsepower of Water Jet (kW)


P Pressure of Water Jet (Bar)
Q Flow Rate of Water Jet (liters/min)

(Note: In English Units: Horsepower = [Pressure (psi) × Flowrate (gpm)]/1,714)

Flow Rate is Function of Nozzle Diameter & Pressure


The flow rate is controlled by the pressure of the feed line and diameter of the
nozzle which can be expressed as:

𝑄𝑄 = 𝑐𝑐1 • 𝑑𝑑 2 • √P (1.2)

Where:

Q Flow Rate of Water Jet (liters/min)


𝑐𝑐1 Empirically determined parameter indicating the efficiency of the water jet. A
greater 𝑐𝑐1 represents less energy loss when the water jet exits from the nozzle.
𝑑𝑑 Nozzle diameter (in millimeters)
P Pressure of Water Jet (Bar)

Water Jet Velocity is Function of Pressure


In turn, the velocity of the water jet exiting the nozzle is related to the water jet
pressure in the feed line which can be expressed as:

𝑉𝑉0 = 𝑐𝑐2 • √P (1.3)

Where:

𝑉𝑉0 Exit velocity of the water jet (m/sec)


𝑐𝑐2 Empirically determined parameter
P Pressure of Water Jet (Bar)

Reaction Force is Function of Pressure and Flow Rate


The reaction force induced by the water jet is also a combined effect of the
pressure and flow rate, expressed as:

40
𝑅𝑅 = 𝑐𝑐3 • √P • Q (1.4)

Where:

𝑅𝑅 Reaction Force (Kg)


𝑐𝑐3 Empirically determined parameter.
P Pressure of Water Jet (Bar)
Q Flow Rate of Water Jet (liters/min)

Specific Energy is Key Metric for Measuring Efficiency


A further useful measure of the relative efficiency of a water jet is Specific Energy.
This is defined as the amount of energy required to remove a unit volume of material from
the work surface, or more specifically:

Specific Energy = Energy/Volume Removed (1.5)

Where Energy is defined as:

𝐸𝐸𝐸𝐸𝐸𝐸𝐸𝐸𝐸𝐸𝐸𝐸 = 𝑐𝑐4 • P • 𝑄𝑄 (1.6)

Substituting from Equation (1.2) for the flow rate, Q, yields:

𝐸𝐸𝐸𝐸𝐸𝐸𝐸𝐸𝐸𝐸𝐸𝐸 = 𝑐𝑐4 • P • 𝑐𝑐1 • 𝑑𝑑 2 • √P (1.7)

Eq. (1.7) can be rearranged and written as:

𝐸𝐸𝐸𝐸𝐸𝐸𝐸𝐸𝐸𝐸𝐸𝐸 = 𝑐𝑐4𝑋𝑋1 • P • √P • 𝑑𝑑 2 (1.8)

Where:

𝑐𝑐4𝑋𝑋1 Product of empirically determined parameters 𝑐𝑐1 and 𝑐𝑐4 .


P Pressure of Water Jet (Bar)
𝑑𝑑 Nozzle diameter (in millimeters)

To determine Volume Removed, we start with the simple expression to determine


the volume removed per second which is simply the product of the slot depth and width
and distance cut per second, or:

41
𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉 𝑅𝑅𝑅𝑅𝑅𝑅𝑅𝑅𝑅𝑅𝑅𝑅𝑅𝑅 𝑝𝑝𝑝𝑝𝑝𝑝 𝑆𝑆𝑆𝑆𝑆𝑆. = Depth • Width • Distance Cut (1.9)

We know from prior studies (Summers 1995) that the depth of cut varies
approximately linearly with pressure and to the 1.5 power with nozzle diameter which,
when substituted into Eq. (1.9) yields the following equation:

𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉 𝑅𝑅𝑅𝑅𝑅𝑅𝑅𝑅𝑅𝑅𝑅𝑅𝑅𝑅 = 𝑐𝑐5 • P • 𝑑𝑑1.5 • 𝑑𝑑 • 𝑣𝑣 (1.10)

Where:

𝑐𝑐5 Empirically determined parameter


P Pressure of Water Jet (Bar)
𝑑𝑑 Nozzle diameter (in millimeters)
𝑣𝑣 Traverse speed with which the nozzle moves over the surface (in millimeters per
sec.)

As Specific Energy is equivalent to the Energy per Volume Removed we divide


Eq. (1.8) by Eq. (1.10) yielding:

𝑆𝑆𝑆𝑆𝑒𝑒𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐 𝐸𝐸𝐸𝐸𝐸𝐸𝐸𝐸𝐸𝐸𝐸𝐸 = (𝑐𝑐4𝑋𝑋1 • P • √P • 𝑑𝑑2 )/( 𝑐𝑐5 • P • 𝑑𝑑1.5 • 𝑑𝑑 • 𝑣𝑣) (1.11)

Eq. (1.11) can be simplified to:

𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆 𝐸𝐸𝐸𝐸𝐸𝐸𝐸𝐸𝐸𝐸𝐸𝐸 = (𝑐𝑐4𝑋𝑋1/5 • √P)/( √𝑑𝑑 • 𝑣𝑣) (1.12)

Where:

𝑐𝑐4𝑋𝑋1/5 Product of empirically determined parameters 𝑐𝑐1 and


𝑐𝑐4 divided by empirically determined 𝑐𝑐5

P Pressure of Water Jet (Bar)


𝑑𝑑 Nozzle diameter (in millimeters)
𝑣𝑣 Traverse speed with which the nozzle moves over the surface (in
millimeters per sec.)

Specific Energy Increases with Jet Pressure


Equation (1.12) demonstrates clearly that the Specific Energy of water jet cutting
increases with jet pressure and decreases with increases in nozzle diameter or nozzle
traverse speed.

42
3.5 Cutting Distances

Cutting Effectiveness Diminishes with Stand-Off Distance


Water jets disperse with distance after being emitted from a nozzle. The distance
from the nozzle orifice to the target is typically referred to as the ‘stand-off distance.’ The
further the distance at atmosphere, the more air that becomes entrapped into the jet stream.
Because of this dispersion, the net horsepower (calculated by the product of pressure and
volume) and associated cutting ability of the jet decreases with distance from the nozzle.

Figure 16 – Breakup Pattern of a Water Jet


(Source: www.blog.kmtwaterjet.com)

Cutting Distance Limited to 200 Nozzle Diameters


Work by Leach and Walker (1966), and industry experience, has yielded a useful
‘rule-of-thumb’ that water jets will cut approximately 150 to 200 nozzle diameters from
the outer end of the jet orifice. Beyond this range, there is a rapid decrease in jet pressure
and corresponding cutting force.

Pressure Decreases Exponentially with Stand-Off Distance


Yanaida (1974), Leach and Walker (1966) and Fowkes and Wallace (1968)
measured the pressure drop in a water jet as a function of the distance from the nozzle exit.
According to their analysis, for a non-dimensional standoff distance, X, the pressure drop
is approximated by:

𝑃𝑃
𝑃𝑃0
= 𝑋𝑋𝑐𝑐 • X −0.8, for 𝑋𝑋𝑐𝑐 < 𝑋𝑋 (1.13)

Where:

𝑃𝑃 Total pressure (psi)


𝑃𝑃0 Water jet stagnation pressure (psi)
𝑋𝑋𝑐𝑐 Non-dimensional length of the initial core region of the water jet

43
X Non-dimensional standoff distance

For X>350, the pressure decreases more rapidly.

Figure 17 – Yanaida (1974) – Jet Dispersion Pattern


(Source: www.blog.kmtwaterjet.com)

As shown in Figure 18 below, pressure decreases exponentially with stand-off


distance per Yanaida’s Approximation Equation (1.13).

0.8 Pressure For Non-Dimensional Stand-Off Distance


0.7
0.6
0.5
Pressure

0.4
0.3
0.2
0.1
0
0 0.5 1 1.5 2 2.5 3
Non-Dimensional Stand-Off Distance

Figure 18 – Yanaida (1974) – Dimensionless Pressure Drop With Distance

Below (Figure 19) is a typical graph of pressure drop-off with stand-off distance
for two identical nozzles. The only difference between the two nozzles is that one has a
four inch straight entrance section which reduces turbulence and improves impact pressure.
More discussion of nozzle design considerations is set out in Section 3.6.

44
Figure 19 – Typical Profile of Pressure Loss with Increasing Stand-off Distance
(Summers 1995)

The Depth-of-Cut Also Diminishes Rapidly with Stand-Off Distance


Similar to the pressure drop-off, the depth-of-cut also decreases dramatically with
distance as depicted in Figure 20 below.

Figure 20 – Typical Profile of Depth-of-Cut with Increasing Stand-off Distance


(Summer 1995)

Cutting Distances in Uraniferous Sandstones – 16 Meters Diameter


Jetting tests conducted by the USBM for an HBHM test in the Teapot sandstone in
Natrona County, Wyoming in 1978 resulted in a mining cavity cut to a radial distance of
7.5 meters from the center of the borehole (15 meters, diameter). The cutting jet pressure
varied from 1,700 to 3,000 psi. A cutting jet pressure of 2,500 psi appeared to be the most
practical. A nozzle diameter of approximately 0.50 inches was used, giving a jet discharge

45
of roughly 325 GPM at 2,500 psi. The jet was able to cut through uraniferous sandstone
at a radial distance of 8 meters (16 meters in diameter) from the borehole.

Cutting Distances in Phosphate Deposit – 11 Meters Diameter


HBHM tests in phosphate with a submerged jet were conducted by the USBM in
1980 in St. John’s County, Florida. A 1” diameter jet, in this case operating at 1,800 psi,
proved effective to mine phosphate at a radial distance of 5.5 meters (11 meters in diameter)
from the borehole. In this case, the cavity was fully submerged in water. When the jet was
fitted with an air shroud injecting a tubular air stream around the water stream at 250 psi
and a low volume of 150 scf/m, the cavity radius had increased by 20% from 4.5 to 5.5
meters.

Air Shrouds Enhance Cutting Distance In Submerged Environments


The increased productivity realized through the air shroud is the result of reducing
the density of the medium in which the jet was cutting. The medium inside the air shroud
is less viscous than the water surrounding it, thus the jet loses energy less rapidly in the air
shroud than when it is cutting through water. The combination of even higher air pressured
and higher volume air flows, within a submerged environment, should both extend the
reach and increase the effective net hydraulic horsepower at the face. This is yet to be
confirmed in practice but theory dictates it to be so.

3.6 Fluid Mechanics & Nozzle Considerations


Cutting Effectiveness Greatly Influenced by Impact Pressure and Nozzle Characteristics
The primary cutting mechanism in a HBHM system is the conversion of kinetic
energy of the water jet into impact pressure (also referred to as stagnation pressure), acting
on the rock face. The maximum impact pressure “P” that can be produced (above the
ambient static pressure) is related to the water velocity “𝑉𝑉0” at the nozzle exit as set out in
Equation (1.14) below:

1
P = 2 • ρ • 𝑉𝑉0 2 (1.14)

Where:
P Pressure of Water Jet (N/𝑚𝑚2 )
ρ Mass density of the jet fluid (kg/𝑚𝑚3 )
𝑉𝑉0 Exit velocity of the water jet (m/sec)

Since the fluid velocity decreases with distance from the nozzle, the impact
pressure, P, is at best proportional to the square of nozzle exit velocity, 𝑉𝑉0. While not
explicitly set-out in Eq. (1.14), the characteristics of the nozzle will also greatly determine
the impact pressure of the fluid on the targeted rock face.

46
Physics of Fluid Flow through the Nozzle
The equations of motion (i.e. continuity, momentum, etc.) governing the ideal flow
through an orifice are employed to model fluid flow. Applying Bernoulli’s equation
between any two points (e.g. Points 1 and 2) in the flow across the nozzle, and neglecting
any changes in elevation between the two points, yields the following relationship:

𝑃𝑃1 𝑉𝑉1 2 𝑃𝑃 𝑉𝑉2 2


+ = ρ2 + (1.15)
ρ1 2 2 2

Where:
𝑃𝑃1 Pressure of Fluid (N/𝑚𝑚2 ) at Point 1
ρ1 Mass density of fluid (kg/𝑚𝑚3 ) at Point 1
𝑉𝑉1 2 Velocity of the fluid (m/sec) at Point 1
𝑃𝑃2 Pressure of Fluid (N/𝑚𝑚2 ) at Point 2
ρ2 Mass density of fluid (kg/𝑚𝑚3 ) at Point 2
𝑉𝑉2 2 Velocity of fluid (m/sec) at Point 2

The continuity equation provides a second relationship between the fluid velocities at
Points 1 and 2 as follows:

ρ1 • 𝑉𝑉1 • 𝐴𝐴1 = ρ2 • 𝑉𝑉2 • 𝐴𝐴2 (1.16)

Where:
ρ1 Mass density of fluid (kg/𝑚𝑚3 ) at Point 1
𝑉𝑉1 Velocity of the fluid (m/sec) at Point 1
𝐴𝐴1 Cross-sectional area (𝑚𝑚2 ) at Point 1

ρ2 Mass density of fluid (kg/𝑚𝑚3 ) at Point 2


𝑉𝑉2 Velocity of the fluid (m/sec) at Point 2
𝐴𝐴2 Cross-sectional area (𝑚𝑚2 ) at Point 2

If the flow areas are circular, as found in pipes, then 𝑉𝑉2 can be expressed as follows:

𝑉𝑉2 2 = 𝑉𝑉1 2 + 2 • (𝑃𝑃1 − 𝑃𝑃2 )/ ρ (1.17)

If ρ1 = ρ2 , and since

𝑑𝑑2
A= π• 4 (1.18)

47
Then

𝑉𝑉1 • 𝑑𝑑1 2 = 𝑉𝑉2 • 𝑑𝑑2 2 , or (1.19)

𝑑𝑑 2
𝑉𝑉1 = 𝑉𝑉2 • (𝑑𝑑2 2 ) (1.20)
1

Therefore, substituting Eq. (1.20) into Eq. (1.17), yields:

𝑑𝑑
𝑉𝑉2 2 = 𝑉𝑉2 2 • (𝑑𝑑2 )4 + 2 • (𝑃𝑃1 − 𝑃𝑃2 )/ ρ (1.21)
1

Simplifying Eq. (1.21) yields:

2(𝑃𝑃 −𝑃𝑃 )
𝑉𝑉2 = � 1 𝑑𝑑224 (1.22)
ρ[1−( ) ]
𝑑𝑑1

Calculating Exit Velocity in English and Metric Units


If 𝑉𝑉2 is the exit velocity of the fluid from the nozzle, then Eq. (1.22) provides an
explicit relationship between pressure drop across the nozzle and the riser pipe diameter.
The following equation provides the water jet exit velocity in English units assuming
𝑙𝑙𝑙𝑙 𝑘𝑘𝑘𝑘
water is the working fluid and ρ = 62.4 𝑓𝑓𝑓𝑓 3, or 1,000 𝑚𝑚3 .

2(𝑃𝑃 −𝑃𝑃 )
𝑉𝑉2 = 12.19 • � 1𝑑𝑑2 24 (1.23)
[1−( ) ]
𝑑𝑑1

Where P is in psi and V is in ft/sec.

2(𝑃𝑃 −𝑃𝑃 )
𝑉𝑉2 = 44.72 • � 1𝑑𝑑2 24 (1.24)
[1−( ) ]
𝑑𝑑1

Where P is in MPa and V is in m/sec.

Calculating Volumetric Flow in English & Metric Units


With the water jet velocity known, the volumetric flow rate can be established
since

Q = 𝑉𝑉 • A (1.25)

48
Where:
Q Flow Rate (liters per minute or U.S. gallons per minute)
V Exit velocity of the water jet (m/sec or ft/sec)
A Cross-Sectional Area of Nozzle Orifice (𝑚𝑚𝑚𝑚2 𝑜𝑜𝑜𝑜𝑖𝑖𝑖𝑖𝑖𝑖ℎ𝑒𝑒𝑒𝑒 2 )

As set out in Eq. (1.25), since 𝑉𝑉2 is the nozzle exit velocity, multiplying by the
cross-sectional area yields the exit volumetric flow. The following equations provide the
flow rate in English and Metric units.

(𝑃𝑃1 −𝑃𝑃2 )
Q = 29.85 • 𝑑𝑑2 2 • � 𝑑𝑑 (1.26)
[1−( 2 )4 ]
𝑑𝑑1

Where P is in psi and, d is in inches and Q is in U.S. Gallons per Minute (GPM).

(𝑃𝑃1 −𝑃𝑃2 )
Q = 2.68 • 𝑑𝑑2 2 • � 𝑑𝑑 (1.27)
[1−( 2 )4 ]
𝑑𝑑1

Where P is in MPa and, d is in mm and Q is in Liters per Minute (LPM).

Coefficient of Discharge Accounts for Nozzle Imperfections


The above equations for flow rate assume a discharge coefficient of 1.0. The
discharge coefficient is an experimental factor which accounts for the imperfections and
associated frictional losses within the nozzle (e.g. turbulence, flow cross-sectional area of
the nozzle not being equal to the actual cross-sectional area, etc.). To account for these
inefficiencies, the actual flow rate, 𝑄𝑄𝑎𝑎 , is the theoretical flow rate (i.e. 𝑄𝑄𝑇𝑇 ) multiplied by
the empirically determined coefficient of discharge for the specific nozzle employed, or

𝑄𝑄𝑎𝑎 = 𝑄𝑄𝑇𝑇 • 𝐶𝐶𝐷𝐷 (1.28)

Where:
𝑄𝑄𝑎𝑎 Actual Flow Rate (liters per minute or U.S. gallons per minute)
𝑄𝑄𝑇𝑇 Theoretical Flow Rate (liters per minute or U.S. gallons per minute)
𝐶𝐶𝐷𝐷 Coefficient of Discharge of Specific Nozzle

Coefficient of Discharge is Experimentally Determined


In practice, the coefficient of discharge is experimentally determined (or provided
by the nozzle manufacturer) and is always less than one. In the event its value is not known,
it can be estimated by measuring 𝑄𝑄𝑎𝑎 . This is important as the estimation of power

49
requirements depends on the magnitude of 𝐶𝐶𝐷𝐷 . The nozzle profile and other characteristics
(e.g. internal surface finish, turbulence flow, etc.) has a significant impact on the value of
𝐶𝐶𝐷𝐷 . The value of 𝐶𝐶𝐷𝐷 approaches 1.0 for a high quality nozzle.

Nozzle Wear Can be Monitored by Measuring Flow Rate over Time


Equations (1.26) – (1.28) are useful since for a given pump, with a maximum flow
rate and pressure, a maximum nozzle size can be determined. They also can be used to
monitor nozzle wear since reductions in pressure are reflected as increases in nozzle
diameter for a given flow rate.

With Flow Rate and Pressure Known, Hydraulic Power Can Be Calculated
With the flow rate and pressure known, Eq. (1.1) can be used to calculate the
hydraulic horsepower (‘E’) of a nozzle, i.e.:

E = P • 𝑄𝑄𝑎𝑎 /1714 (1.29)

Where:
E Horsepower
P Pressure of Water Jet (psi)
𝑄𝑄𝑎𝑎 Actual Flow Rate (U.S. Gallons per Minute)

Or,

E = 16.66 • P • 𝑄𝑄𝑎𝑎 (1.29)

Where:
E Watts
P Pressure of Water Jet (MPa)
𝑄𝑄𝑎𝑎 Actual Flow Rate (Liters per Minute)

As set out in Equations (1.1), (1.28) and (1.29), the hydraulic power of the water jet at the
exit of the nozzle is the product of the pressure and flow rate.

Hydraulic Power is Function of Nozzle Diameter and Pressure


If the previous Eqs. (1.26) - (1.28) are substituted into Eqs. (1.1), (1.28) and (1.29),
the power becomes a function of pressure and nozzle size as expressed below:

3
E = 𝑑𝑑2 •𝑃𝑃 2 (1.30)

50
Where:
E Watts
𝑑𝑑 Nozzle Diameter (mm)
P Pressure of Water Jet (MPa)

Hydraulic Power Varies as Square of Nozzle Diameter & Three-Halves Power of Pressure
As set out in Equation (1.30), the hydraulic power of the water jet exiting the nozzle
varies as the square of the nozzle diameter and to the three-halves power of the pressure.
In other words, power output is more sensitive to changes in nozzle diameter than pressure.
(See example in paragraphs below). This relationship underscores the importance of
minimizing and controlling the nozzle diameter of the jet. Doubling the nozzle diameter
increases the water jet’s hydraulic power by a factor of four while doubling the pressure
increases the power by a factor of 2.8.

The Importance of Nozzle Size


The pumps used in HBHM systems typically operate on a positive displacement
basis which means that a fixed volume of water leaves the pump and enters the delivery
line to the nozzle each minute. When all other outlets are closed, this volume must pass
through the orifice in the nozzle body. The jet acquires its final velocity as a result of
having to pass through the small hole in the nozzle. The pressure from the pump is that
which is required to drive the total volume of water through the nozzle orifice. Thus, with
a pump which puts out a steady flow of water, the pressure of the water jet exiting the
nozzle orifice is controlled by the size of the nozzle orifice.

A Real Life Example


Where the flow rate is 20 lpm and the required pressure is 350 bar, we can calculate
the size of the required orifice as follows:

One liter of water occupies 1,000 cc. Thus 20 liters will occupy 20,000 cc. If this
must pass through the orifice in one minute, then one-sixtieth of this, or 333 cc, much pass
through each second. In an ideal world, the velocity which a given pressure will generate
is given by Equation (1.3), namely:

𝑉𝑉0 = 𝑐𝑐2 • √P (1.3)

Where:

𝑉𝑉0 Exit velocity of the water jet (m/sec)


𝑐𝑐2 Empirically determined parameter indicating the efficiency of the water jet
P Pressure of Water Jet (Bar)

51
Using a 𝑐𝑐2 that has been empirically determined to be 17.14 (Summers 1995), and
substituting a jet pressure of 350 bar into Equation 1.3 yields the jet velocity, 𝑉𝑉0, as follows:

m
𝑉𝑉0 = 17.14 • √350 = 321 s
(1.31)

The jet must thus move through the selected orifice at 321 m/s or 32,100 cm/sec.
Since we much move 333 cc each second this implies that the orifice must be 333/32,100
= 0.01037 sq. cm. From the simple equation for the area of a circle:

𝜋𝜋
𝐴𝐴𝐴𝐴𝐴𝐴𝐴𝐴 = �4 � • 𝑑𝑑 2 , (1.18)

this is equivalent to a diameter of 1.15 mm.

Small Changes in Nozzle Size Greatly Impact Velocity & Jet Pressure

To sum up the foregoing, to get a pressure of 350 bar from a pump which is putting
out 20 lpm of water we would need to drive that water through a hole some 1.15 mm in
diameter. Small changes in nozzle size can have a significant effect on the jet pressure as
demonstrated in the Table 3 and Figure 21 below (Summers 1995). A modest increase in
nozzle diameter size from 1.00 to 1.15 mm, at a constant volume flow of 20 lpm, reduces
fluid velocity from 424 to 321 (m/sec) and jet pressure from 613 to 350 bar.

Nozzle Nozzle Volume Fluid Jet


Diameter Area Flow Velocity Pressure
(mm) (sq. cm.) (lpm) (m/sec) (bar)
1.00 0.00785 20 424 613
1.05 0.00865 20 385 504
1.10 0.00950 20 351 419
1.15 0.01038 20 321 350
1.00 0.00785 21 446 675
1.00 0.00785 22 467 741
1.00 0.00785 23 488 810

Table 3 – Relative Jet Performance for Varying Nozzle Sizes


(Summers 1995)

52
Jet Pressure vs. Nozzle Diameter
650

600

550
JET PRESSURE (BAR)

500

450

400

350

300
0.95 1.00 1.05 1.10 1.15 1.20
NOZZLE DIAMETER (MM)

Figure 21 – Jet Pressure (Bar) vs. Flow Rate (lpm)

Changes in Flow Rate Also Greatly Impact Jet Pressure


Flow rate also has a significant effect on jet pressure. (See Figure 22 below).
Increasing the flow rate from 20 to 23 lpm (i.e. 15%) will increase the pump pressure
required to deliver that volume through the same 1.00 mm diameter orifice from 613 to
810 bar, or approximately 30%.

Jet Pressure vs. Flow Rate


850

800
JET PRESSURE (BAR)

750

700

650

600
19.5 20 20.5 21 21.5 22 22.5 23 23.5
FLOW RATE (LPM)

Figure 22 – Jet Pressure (Bar) vs. Flow Rate (lpm)

53
The foregoing highlights the sensitivity of nozzle design parameters and their
critical influence on the overall performance of an HBHM system.

A Real Life Complication – Coefficient of Discharge


The above calculations are somewhat idealized and ignore real world
complications. The most important complication is the underlying assumption that the
water jet will discharge from the nozzle at the same size of the orifice. In actual fact, the
water will converge to a smaller size than the orifice either inside or just outside the nozzle
body. This results in a smaller area and thus a faster discharge velocity for the water. To
take into account this effect, one needs to multiply the area of the nozzle by a Coefficient
of Discharge. This can vary from about 0.60 for a poor nozzle design to approximately
0.95 for a very good nozzle design. Labus (1989) has documented the calculations and
equations which inter-relate the flow rate through a system and the pressure and pipe and
jet diameters. He further provides example Coefficients of Discharge for varying generic
nozzle shapes. His analysis is beyond the scope of this thesis.

Nozzle Design is a Critical Component in An HBHM System


Notwithstanding the fact that nozzle design and performance is a critical component
of an HBHM system, it has received scant attention from past researchers and in historic
pilot tests. Considerably greater industry and academic attention has been focused on
nozzle design for cutting and cleaning applications involving high pressures (70,000 psi)
and low flow rates (< 25 gpm) where the nozzle can be held relatively close to the targeted
material. Such designs are inadequate for HBHM applications which require larger
nozzles, lower pressures (< 5,000 psi), higher flow rates (> 250 gpm) and a longer standoff
distance to achieve economic volumes of cutting. An additional technical challenge unique
to HBHM water jet nozzle design is the requirement to redirect the water flow 90 degrees
from the borehole conduit to the nozzle inlet. In making this turn, the water flow will
develop a swirling turbulence component which will severely deteriorate the coherency of
the water jet if not properly accounted for in the nozzle design.

54
Figure 23 – Water Jet Nozzle Design Parameters
(Source: www.blog.kmtwaterjet.com)

Past Studies
The first major study of water jet nozzle design was carried out by Rouse et. al.
(1952) to improve the performance of fire fighting monitors. This was followed by a
definitive study by Leach and Walker (1966) who evaluated rock cutting effectiveness
across a spectrum of nozzle designs. Their study was mainly concerned with the effect of
nozzle shape on the reduction in velocity caused by the passage of the water jet through
air. As their study demonstrated, the nozzle design has an important effect on the
disintegration of the jet and hence on the speed of the jet at some distance from the nozzle.
The Leach and Walker study concluded that the most effective shape was that proposed in
an earlier study by Nikonov and Shavlovskii (1961). The simplicity of the Nikonov design
lies in the two parts of the nozzle. The first of these being the fluid acceleration section
with the channel linearly narrowed at a 13 degree included angle from the pipe diameter to
the orifice size. The second section, the straight section or ‘throat’, is fixed at 2.5 times
the nozzle diameter. Figure 24 below illustrates the classic Nikonov nozzle design.

Figure 24 – Nikonov Nozzle Design


(Source: www.blog.kmtwaterjet.com)

55
Subsequent work in Russia suggests that the conic angle should be decreased as the
jet velocity increases, and conversely, can be increased to 18 to 20 degrees at lower jet
pressures for optimum water jet performance (Summers, 1995). A variety of nozzle designs
is depicted in Figure 25 below.

Figure 25 – Variety of Water Jet Nozzle Designs


(Source: www.blog.kmtwaterjet.com)

Flow Straighteners – A Necessary Feature of HBHM Nozzle Designs


In HBHM applications, the water jet flow must be turned 90 degrees from the
borehole conduit to the nozzle inlet. The turning of the fluid has the potential to generate
severe turbulence and spin. If the flow retains a high degree of turbulence as it passes
through the nozzle then the jet structure beyond the nozzle, and cutting effectiveness, will
be sub-optimal. Older studies have suggested that the pipe immediately preceding the
nozzle be straight and aligned with the nozzle axis for at least 100 nozzle diameters to
minimize turbulence flow into the nozzle. This design luxury is not possible in HBHM
systems due to the small borehole diameter size. Subsequent work by researchers
introduced a ‘flow straightener’ behind the nozzle section which yields acceptable results
in much shorter distances than the idealized straight section of 100 nozzle diameters.
Various flow straightener geometries and configurations have been introduced and shown
in practice to be effective in minimizing turbulence. (See Figure 26 below for example
geometries for flow straighteners). Shavlovskii (1972) has shown that a flow straightener
can reduce the length of the straight section from an ideal ‘100’ to ‘10’ nozzle diameters.
Within the straightener section, he recommends that the linear plates of the stabilizer
should be 2-4 times the diameter of the inner pipe. He further recommends that there
should be a gap between the flow straightening device and nozzle inlet of less than two
pipe diameters.

56
Figure 26 – Sample Flow Straightener Geometries (Summers 1995)

Good Nozzles vs. Bad Nozzles


Nozzles of the same shape and size do not necessarily give the same jet
performance. Subtle internal design considerations can result in considerably different
performance outcomes. As a general proposition, design choices must serve to minimize
turbulence. Straight edges and sharp transitions that give rise to fluid spin are to be
avoided. In the case of multiple orifices, it is beneficial to divide the central fluid flow
prior to acceleration into individual nozzle inlets. (Figure 27 below from Summers (1995)
illustrates a “bad” vs. “good” internal nozzle design for a two jet system. Note the splitting
of the central fluid flow prior to acceleration into individual nozzle inlets.

Figure 27 – Comparison of Bad (‘a’) vs. Good (‘b’) Nozzle Design for Two Jet System (Source:
www.blog.kmtwaterjet.com)

Finish, Flow Straightness and Shape are Key


The most important parameters for internal nozzle design are:
• internal surface finish of the nozzle;

57
• straightness of the flow into the orifice; and
• shaping of internal nozzle transitions, ‘tapered’ being preferable to ‘sharp’

While the effective cutting distance in practice is generally limited to a maximum of


200 nozzle diameters from the orifice, it can be much less for a poorly constructed nozzle
of the same dimensions. Summers (1995) presents a compelling illustration of the jet
performance for ‘Bad’ (Figure 28) vs. ‘Good’ (Figure 29) nozzle designs of identical
dimensions for a two jet system.

Figure 28 –‘Bad’ Nozzle Design - Pressure Profiles for Two Jet System (Summers 1995)

Figure 29 –‘Good’ Nozzle Design - Pressure Profiles for Two Jet System
(Summers 1995)

58
Nozzle Design for HBHM Applications – Lohn and Brent Study
To address the unique requirements of HBHM applications, the USBM funded a
study of optimal nozzle design for mineral excavation (Lohn and Brent, 1976). The
objective of the Lohn and Brent study was to propose improved nozzle designs for HBHM
applications where flow rates and standoff distances are large and the water flow must be
redirected 90 degrees from the borehole conduit to the nozzle inlet. Specific design
considerations in their study were as follows:
• Borehole radius is 9 inches, thereby allowing only 4 inches for flow turning, 4
inches for nozzle contraction and 1 inch for clearance;
• A pre-nozzle velocity of 20 feet per sec. is assumed; and
• A water flow rate of 200+ gpm is required and a pressure of 1,500+psi.

Conclusions from Lohn and Brent Model – Turbulence Matters


Lohn and Brent developed a series of differential equations to model the fluid dynamics
of water flow through an idealized nozzle design. They then developed a numerical
solution for their system of differential equations which yielded the following general
conclusions:
• Large curvature at the nozzle entrance is not desirable because it can cause
boundary layer separation.
• Large curvature at the nozzle exit is not desirable because it can cause cavitation.
• The length of the nozzle should be kept small to increase the acceleration and
decrease the boundary layer thickness at the nozzle exit. The shortness of the
nozzle is limited by separation/cavitation considerations.
• Design consideration should give priority to the jet diameter rather than the nozzle
diameter.

Quartic-Straight Nozzle Design Recommended for HBHM Applications


Lohn and Brent recommended a quartic-straight design for HBHM applications
with an exit half angle of 20 degrees. (See Figure 30 Below). They cautioned that special
design attention is required in redirecting the water flow 90 degrees from the borehole
conduit to the nozzle inlet to minimize the swirling component and introduced a flow
curving and straightening section in their final design. Finally their analysis suggested that
most HBHM applications will require a pressure of at least 2,000 psi and flow rate of 200+
gpm to optimize cutting effectiveness.

59
Figure 30 – Lohn and Brent HBHM Nozzle Design
(Source: www.blog.kmtwaterjet.com)

Lohn and Brent Design Found Effective in HBHM Pilot Tests


The Lohn and Brent design was incorporated in a subsequent USBM designed
HBHM tool and found to be effective in a number of pilot tests involving uranium, coal
and phosphate ores (See Section 3.1). Given the limited industry use and testing of HBHM-
specific nozzles, further design improvements are likely.

3.7 Importance of Pressure and Flow Rates

Cutting Rock Matrix Due to Power in Water Jet


Cutting rock matrix or mining with water jets is due to power in a water stream and
the transmission of this power to a rock thereby causing fragmentation. Power is imparted
to the water by a pump which pushes a given volume of water per unit time into a feed line
while storing pressure energy in it. This pressurized water flows through the feed line to a
jet nozzle which has an orifice with a diameter much smaller than that of the feed line.
Since the constant flow of water reaches the nozzle, the water must accelerate to a high
velocity to escape through the orifice. The water is thereby shaped into a coherent stream,
or jet, of high velocity water.

Turbulence Control is Critical


Control of the flow of water is critical to maximizing cutting efficiency. Extensive
work has been completed over the past fifty years in determining the optimum jet size and
dimension. Minimization of turbulence within the system is of paramount importance and
critical to the design of any HBHM system. Turbulence can enter the system at many
points starting with the pumping system, at the ninety degree turn that is required to re-
orient the jet from the borehole to the rock face, within the nozzle as the water flow
increases velocity prior to nozzle exit, in lines, pipe fittings, valves, etc. System

60
components need to be carefully selected with turbulence minimization in mind. Because
of line losses, turbulence and other physical and mechanical inefficiencies it is not unusual
in practice for up to 85% of the input energy to be dissipated prior to the water reaching its
target (See Section 3.9).

Water Jet Kinetic Energy Converted to Impact Pressure on Rock Face


The moving stream of water can do work only when it impacts the rock face. Once
the water exits the nozzle it is no longer under pressure but has stored energy in the form
of kinetic energy. For a pump to push a given volume of water through a feed line and
nozzle, it must exert a pressure on the water. This pressure is expended in two ways; by
driving the water through the feed line from the pump to the nozzle and pushing the water
through the nozzle at an increased velocity. When the water jet reaches the rock face, the
kinetic energy is converted into an impact pressure which penetrates and fragments the
rock.

3.8 Rock Penetration

Rock Penetration Dependent on Tool Parameters and Ore Type


The cutting of a specific rock type is dependent on a multitude of factors involving
both the physical properties of the rock and tool operating parameters. Hardness, surface
abrasion resistance and the density of the rock are the most important rock properties for
assessing cutting effectiveness. From the tool side, flow rate and water velocity are the
key parameters that determine the effectiveness of the water jet in fragmenting the rock
face. The first being the volume of water hitting the rock face (i.e. flow rate) and the second
being the water velocity (i.e. the impact pressure). The impact pressure breaks the rock
and the fragments are transported away from the rock face by the water flow. The power
of the water jet is the product of the pressure and flow rate, i.e. Eq. (1.1). The most efficient
use of the energy in the jet is attained by selecting that combination of pressure and flow
rate which fragments and carries away the greatest volume of rock per unit time. This
combination will effectively deliver sufficient water flow and pressure to cause significant
impact at the rock face to cause it to fragment into a slurry with solids small enough to
return up the airlift tract and be recovered through the HBHM equipment at surface.

Farmer and Attewell’s Rock Penetration Model


In Farmer and Attewell’s (1965) seminal study of rock penetration by water jet,
they used the concept of momentum transfer to predict the penetration of a jet into a rock
target. They concluded that the depth of penetration (or cut) was directly proportional to
the momentum of the projectile as given by the following equation:

𝐾𝐾
𝑆𝑆 = �𝐴𝐴 � • 𝑚𝑚 • (𝑣𝑣0 − 𝑣𝑣𝑐𝑐 ) (1.32)

61
Where:

𝑆𝑆 Depth of Penetration
𝑘𝑘 Empirical constant depending on the mechanical properties of the target material
and of the projectile
𝐴𝐴 Cross –sectional areas of the projectile
𝑚𝑚 Mass of the projectile
𝑣𝑣0 Terminal impact velocity
𝑣𝑣𝑐𝑐 Critical velocity at which penetration first occurs

Volume Matters
The shape of the crater also has an important bearing on its volume. If the crater is
cylindrical, its volume, VOL, is proportional to S, and hence to the impact velocity, 𝑣𝑣0 , for
the same projectile constants. However at higher velocities, and in hard brittle materials
where impact leads to irregular breakage and a conical or bell shaped crater, VOL is better
defined as proportional to the kinetic energy of impact, i.e.

𝑚𝑚•𝑣𝑣0 2
𝑉𝑉𝑉𝑉𝑉𝑉 = 𝑘𝑘1 • � � (1.33)
2

Where:

𝑉𝑉𝑉𝑉𝑉𝑉 Volume of Penetration


𝑘𝑘1 Empirical constant depending on the mechanical properties of the target material
and of the projectile
𝑚𝑚 Mass of the projectile
𝑣𝑣0 Terminal impact velocity

Farmer and Attewell Study Results Helpful But Limited


The Farmer and Attewell study highlighted the importance of nozzle design in
projecting a cohesive jet of water to maximize cutting effectiveness. It also emphasized
the critical need to minimize turbulence in the nozzle. Their experimental studies
suggested that for maximum penetration into softer rocks a nozzle diameter of less than 3
mm is desirable and for harder rocks a still small nozzle diameter is preferred. For mining
applications, they suggested a pump capable of delivering up to 50 lpm at pressure up to
7000 kg/sq. cm would be desirable. Although the Farmer and Attewell study gave some
indication of the relative performance of a jet on a target, it did not include any
consideration of the target response as a means predicting the actual penetration depth.
Additional theoretical work was therefore required. It was Rehbinder who provided this
critical follow-up work.

62
Rehbinder’s Theory – How Water Jets Cut Rock – A Big Step Forward
In a series of papers, Rehbinder (1977, 1978, 1979) proposed a theoretical model
on a micro-scale to describe how a water jet cuts rock. The underlying idea is that if a
water jet hits the surface of a rock the water starts to penetrate the pores between the grains.
If the jet traverses over the surface the penetration ceases once it passes. This implies that
the penetration takes place during the time of exposure T= d/v, where d and v are the
diameter and traversing velocity of the jet, respectively. Rehbinder considers two cases.
In the first case the stagnation pressure of the jet 𝑝𝑝0 is less than the threshold pressure of
the rock 𝑝𝑝𝑡𝑡ℎ which implies that the jet causes no damage to the rock face. In the second
case the stagnation pressure of the jet is greater than the threshold pressure (𝑝𝑝𝑡𝑡ℎ ) of the
rock (i.e. 𝑝𝑝0 >𝑝𝑝𝑡𝑡ℎ ) which implies that the grains break away from the rock face at a rate
equal to the mean rate at which the water passes a grain. The velocity ṽ is assumed to be
given by Darcy’s law for flow through a porous media, i.e.:

𝑥𝑥𝑃𝑃
ṽ= • 𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔 𝑝𝑝 (1.34)
𝑢𝑢

Where:

ṽ Velocity of grain disaggregation


𝑥𝑥𝑃𝑃 Modified permeability of the rock
𝑢𝑢 Dynamic viscosity of water
𝑝𝑝 Pressure of water in the pores

Hydraulic Lifting Force Does The Cutting


If 𝑝𝑝0 <𝑝𝑝𝑡𝑡ℎ , no erosion takes place but nevertheless the water penetrates the rock
to a certain depth given by Darcy’s law and the time of exposure. If, however, 𝑝𝑝0 >𝑝𝑝𝑡𝑡ℎ , the
grains are not only passed by the penetrating water but also removed grain-by-grain by
hydraulic lift forces which Rehbinder defines as follows:

V
𝐹𝐹 = (1−𝒫𝒫) • 𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔 𝑝𝑝 (1.35)

Where:

𝐹𝐹 Hydraulic force that acts against a grain of rock exposed to a water jet due to
viscous drag.
V Volume of the grain of rock
𝒫𝒫 Porosity of the rock grain
𝑝𝑝 Pressure of water in the pores

63
Figure 31 – Illustration of Rehbinder (1979) Water Droplet Hydraulic Lift Force (Source:
www.blog.kmtwaterjet.com)

If Ore Impermeable, No Cutting Takes Place


Rehbinder’s Equation (1.35) is otherwise known as the Threshold Pressure and
implies that if the water is not allowed to enter the rock there will be no viscous drag acting
at the grains and no damage. This somewhat surprising result was confirmed
experimentally by Rehbinder. Another interesting consequence of Equation (1.35) is that
a coarse grained rock is easier to penetrate and cut than a fine grained specimen of the same
type since the force is proportional to the volume of the grain.

Rate and Geometry of Cut


The rate at which a slot or cut would be created, provided that the water jet was
given sufficient time to penetrate the surface, was determined by Rehbinder to be:

𝑑𝑑ɦ
𝑑𝑑𝑑𝑑
= ṽ, (1.36)

where ɦ is slot depth and ṽ is the velocity of grain disaggregation as set out in Equation
(1.34). Equation (1.35) requires the pressure at the depth of the slot that Rehbinder has
proposed to be given by the relationship:

𝛽𝛽ɦ
𝑝𝑝
𝑝𝑝0
= 𝑒𝑒 − 𝐷𝐷 (1.37)

Where:

𝑝𝑝 Pressure at depth of the slot


𝑝𝑝0 Water jet stagnation pressure

64
𝛽𝛽 Empirical constant for a specific rock type
ɦ Slot depth
𝐷𝐷 Width of the slot

Equations (1.34) thru (1.37) can be used to solve a series of differential equations to yield
the relationship that so long as the pressure generated at the bottom of the slot exceeds the
threshold pressure (𝑝𝑝𝑡𝑡ℎ ), then the slot geometry will be defined by:

ɦ 1 𝛽𝛽•𝑥𝑥𝑃𝑃 •𝑝𝑝0
𝐷𝐷
= 𝑑𝑑𝑑𝑑 • log 𝑛𝑛 (1 + • 𝑇𝑇) (1.38)
𝑢𝑢•ḹ•𝐷𝐷

Where:

ɦ Slot depth
𝐷𝐷 Width of the slot
𝛽𝛽 Empirical constant for a specific rock type
𝑥𝑥𝑃𝑃 Modified permeability of the rock
𝑝𝑝0 Water jet stagnation pressure
𝑢𝑢 Dynamic viscosity of water
ḹ Average grain diameter
𝑝𝑝 Pressure of water in the pores
T Time of exposure T= d/v, where d is the jet diameter and v is the traverse velocity.

This in turn defines the maximum depth to which the slot can be cut as:

ɦ 1 𝑝𝑝
max(𝐷𝐷) = 𝛽𝛽 • log 𝑛𝑛 (𝑝𝑝 0 ) (1.39)
𝑡𝑡ℎ

Where:

ɦ Slot depth
𝐷𝐷 Width of the slot
𝛽𝛽 Empirical constant for a specific rock type
𝑝𝑝0 Water jet stagnation pressure
𝑝𝑝𝑡𝑡ℎ Threshold pressure

Cut Shapes Vary with Stand-Off Distance and Traverse Rate


Figure 32 below illustrates the cut shape variation with stand-off distance and
traverse rate.

65
Increasing Stand-Off Distance →

Increasing Traverse Rate↓

Figure 32 – Geometry of Cut Shape Variation with Stand-Off Distance & Traverse Rate
(Summers 1995)

Rock Cutting is Dependent on Threshold Pressure and Erosion Resistance


In Rehbinder’s third paper (1979) he advances his earlier work by developing a
more robust set of predictive cutting models for specific rock types. He postulates that the
two rock parameters which are relevant when a rock is slotted by a water jet are: (i) the
Threshold Pressure (Eq. 1.35); and (ii) the Erosion Resistance which he defines as follows:

1 ḹ
𝑅𝑅𝑠𝑠 ∝ ḹ • (𝛿𝛿)2 (1.40)

Where:

𝑅𝑅𝑠𝑠 Erosion resistance


ḹ Average grain diameter
𝛿𝛿 Average pore diameter

Equation (1.40) implies that erosion resistance is proportional to the inverse of the
grain diameter due to the fact that the advance rate is assumed to take place in steps of
magnitude ḹ, the average grain diameter. This is to say that the rate and depth of cutting
are an inverse function of grain size; i.e. a larger grain rock is easier to cut to a greater
depth than finer grained rocks. Experimental work by Rehbinder, Summers (1995) and
others has confirmed this to be so.

66
Cutting Efficiency Increases with Traverse Speed Up to Limit
From Equations (1.38) and (1.39), Rehbinder predicted that slot cutting efficiency
would increase with traverse speed up to some inherent limit, which was a function of rock
permeability. For permeable rock, this limit would be approximately 100 m/sec. Figure
33 illustrates this for two identical rock types with the only difference being their respective
permeability. As noted in Figure 33, rock permeability has a significant impact on cutting
efficiency suggesting that structural properties are critical to predicting cutting efficiencies.
The subject of physical vs. structural properties is explored in more detail in the next
section (Section 3.9).

Figure 33 – Typical Profiles: Depth Penetration vs. Traverse Rate For Two Rock Permeabilities
(Summers 1995)

Real World More Complex Then Rehbinder’s Equations


Subsequent testing of Rehbinder’s equations in laboratory experiments has yielded
mixed results (Summers 1995) suggesting that nature is more complex than Rehbinder’s
predictive models. Notwithstanding this, Rehbinder’s work has provided a solid
theoretical basis which will require further refinements by future research.
In real-world practice, lab testing will be required to make an initial determination
of the appropriate flow rate and pressure rate settings to achieve the desired rock cutting
volumes and efficiencies. Pilot testing will then be required to make adjustments to these
parameters along with final dimensions and design of the water jet itself prior to the
commencement of commercial scale operations.

67
3.9 Rock Properties – Physical vs. Structural

Historic Research Focused on Physical Rock Properties


The cutting effectiveness of the water jet is a function of both the operating
parameters of the tool (e.g. impact pressure, water volume, traverse rate, etc.) and specific
rock properties (e.g. density, hardness, compressive and tensile strengths, porosity, grain
size, fracture density and distribution, etc.). Various researchers over the past five decades
have attempted to develop analytical and/or empirical models to predict the cutting
effectiveness of water jets and mechanical tools (e.g. cutter heads). Thiruvengadam (1965)
proposed a series of equations defining “erosion strength” as a function of fluid impact
force and the target material’s strength. Heymann (1970) extended this work by
developing an alternative predictive model of erosion resistance based on empirical data
and dimensional considerations. Kinoshita et. al. (1972) found that eroded volume was
inversely related to the compressive strength of the rock. Similarly, in an examination of
twelve physical properties of seven different rock types, Summers (1972) found that the
depth of cut was most closely correlated with two rock properties; i.e. the inverse of
compressive strength and Shore Hardness. Tecen (1982) examined water jet assisted tool
cutting of rocks and hypothesized that the best predictor of cutting effectiveness was rock
compressive strength, NCB Cone Indenter Index, porosity and the density of micro
fractures. Agus et. al. (1993) correlated the specific energy (Eq. 1.5) of water jet cutting
with two principal rock properties, i.e. the point load strength of the rock and the sonic
wave velocity of P-waves propagated through the rock mass. Tiryaki (2006) investigated
the relationship between measures of rock brittleness and fracture toughness on specific
cutting energy for various rock types and concluded that a rock’s NCB Cone Indenter Index
measure was the best predictor of specific cutting energy. Engin (2012) conducted a
comprehensive analysis of abrasive water jet cutting depth for 42 natural stones and found
that Shore Hardness, Bohme Surface Abrasion Resistance and rock density were the most
significant rock properties affecting cutting depth.

Structural Rock Properties Are More Important


While the above noted studies (and many more not cited) suggest various physical
rock properties (especially compressive strength and hardness) as important predictors of
water jet cutting effectiveness, it is easy to cite counter examples which serve to
demonstrate their collective limitations. Summers (1995) did so by cutting two rocks of
approximately equal compressive strength and surface hardness at the same traverse tool
speed and other water jet parameters. The cuts in the finer grained limestone were
considerably smaller than in the larger grained sandstone, suggesting that structural
properties of the rock are perhaps as important as physical properties, if not more so.

68
Fracture Density Likely To Be Controlling Factor
As Rehbinder (1980) has suggested, water jets attack a rock by inducing failure at
the weakest point in the surface and hydraulically lifting individual grains from existing
fractures and cracks in the rock mass. In terms of cutting performance, the controlling
physical parameters of the rock are likely to be fracture density (inversely related to grain
size) and the surface energy of the rock. As most rock masses are heterogeneous and
anisotropic, developing an all encompassing analytical or empirical prediction model of
cutting effectiveness will likely remain an elusive goal. This has been further underscored
by Alehossein et. al. (2004) who demonstrated that even for samples cored from the same
rock outcrop, with uniform geology, geochemistry and mineralogy, variances in fracture
size and distribution, unconfined compressive strength and tensile strength can be
significant.

Griffith Provides Basic Equation for Fracture Growth


Fractures and their propensity to grow and expand under hydraulic pressure are key
to the performance of a water jet. Griffith (1920, 1924) used thermodynamic arguments to
determine necessary conditions for a crack (or fracture, we use the words interchangeably)
to grow due to an applied load. His simplified model comprised a rock specimen of length
L, width h, and thickness t, containing a thin crack of length 2c lying perpendicular to the
side of length L. He imagined that a tensile load T is applied to the rock by hanging mass
m connected to the rock by a cord that passes over a frictionless pulley. Griffith further
assumed his system to be in equilibrium with the crack length having length c. The total
energy of Griffith’s idealized model consists of the elastic strain energy in the rock, the
surface energy of the crack, and the gravitational potential of the mass.
Griffith’s analysis hinges on the insight that, as a crack grows, energy is needed to
create new surface area and this is only thermodynamically possible if extension of the
crack allows the total energy of the system to decrease. In Griffith’s model, the critical
crack length for maximizing the total energy as a function of c is found by fixing all other
variables and taking the derivative of the total energy of the system. The maximum energy
will exist where the derivative is equal to zero and can be found by solving the equation:

𝑑𝑑𝐸𝐸𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡 𝑑𝑑𝐸𝐸𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒 𝑑𝑑𝐸𝐸𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠 𝑑𝑑𝐸𝐸𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔


= + + =0 (1.41)
𝑑𝑑c 𝑑𝑑c 𝑑𝑑c 𝑑𝑑c

Where:
c: Half Length of the Crack
𝐸𝐸𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡 : Total Energy of the System
𝐸𝐸𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒 : Elastic Energy
𝐸𝐸𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠 : Surface Energy
𝐸𝐸𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔𝑔 : Gravitational Energy

69
Through various assumptions and substitutions, Equation (1.41) becomes:

2•π•𝜎𝜎𝑐𝑐 2 •t•c
4•γ•t− 𝛦𝛦
=0 (1.42)

Where:
𝑐𝑐: Crack length
γ: Surface Energy per Unit Length of crack Surface
𝛦𝛦: Young’s Modulus
t: Rock specimen thickness
𝜎𝜎𝑐𝑐 : Tensile Stress

As seen from the above Equation (1.42), the derivative of the total energy of the system is
positive for small values of c and negative for large values of c.

By simplifying and rearranging terms of Equation (1.42), one can solve for the
critical length c ∗ above which cracks are thermodynamically free to grow at fixed tensile
strength stress 𝜎𝜎𝑐𝑐 (see Jaeger, J.C. et. al (2007).

2•γ•𝛦𝛦
c ∗ = π•𝜎𝜎 2 (1.43)
𝑐𝑐

Where:
c∗: Critical crack length above which cracks are thermodynamically free to grow
γ: Surface Energy per Unit Length of crack Surface
𝛦𝛦: Young’s Modulus
𝜎𝜎𝑐𝑐 : Tensile Stress

Bieniawski’s Extension of Griffith’s Model of Fracture Growth


Bieniawski (1967), building on earlier work by Griffith, rearranged terms of
Equation (1.43) to determine the critical tensile stress 𝜎𝜎𝑐𝑐 at which a crack of initial length
c will be able to grow. This yields Equation (1.44), the critical tensile stress 𝜎𝜎𝑐𝑐 which is a
necessary condition that must be satisfied for the crack growth to be thermodynamically
possible.

𝜎𝜎𝑐𝑐 = √(2γ𝛦𝛦/πc) (1.44)

Where:
𝜎𝜎𝑐𝑐 : Applied Uniaxial Tensile Stress At tip of crack
γ: Surface Energy per Unit Length of crack surface
𝛦𝛦: Young’s Modulus

70
c: Half Length of the crack

While Equation (1.44) is a necessary condition that must be satisfied for crack
growth, it is not in itself sufficient to cause growth to occur. The sufficient condition is
that the tensile stress at the tip of the crack must also be large enough to break the atomic
bonds at the tip crack. However, as noted by Jaeger J.C. et. al. (2007), this latter condition
will in practice always be satisfied as the stress concentration at the tip of an elliptical crack
of a small aspect ratio is quite large.

A noteworthy implication of Equations (1.43) and (1.44) is that small cracks (i.e.
small c) will require relatively large stresses to allow crack growth whereas large cracks
will grow under conditions of smaller tensile stress.

The condition given in Eq. (1.44) suggests that for

𝜎𝜎𝑐𝑐 < √(2γ𝛦𝛦/πc) (1.45)

the crack will not extend.

For the condition where the stress at the tip of the crack exceeds the value returned by
Equation (1.44), i.e.

𝜎𝜎𝑐𝑐 > √(2γ𝛦𝛦/πc), (1.46)

the crack will grow as it intersects around individual grains, liberating them and exposing
the underlying fracture(s) to penetration by the water jet.

Griffith’s Model is Useful but has Real World Limitations


Griffith’s seminal model, and its subsequent variations and extensions, provide a
useful framework for understanding crack formation. However, Griffith’s simplistic model
has its limitations in predicting real world behavior of rock failure. Numerous researchers
have attempted to develop alternative, more complex approaches to predicting rock failure
with mixed results. These are beyond the scope of this thesis but Patterson and Wong
(2005) and Yuan and Harrison (2006) have reviewed numerous stochastic approaches that,
in theory, better cope with the complexities of a heterogeneous, anisotropic rock mass.
Most of these approaches involve division of the rock mass into smaller discrete
units/blocks with each being assigned physical parameters and probabilities. Relationships
between the discrete units are also mathematically defined and various solution approaches
are employed to predict rock failure. All these predictive models suffer from challenges in

71
deriving realistic data inputs as well as overwhelming mathematically and computational
complexity as the size of the data units increases.

The Presence of Fluids Adds Complexity


The presence of fluids from the water jet adds significant complexity to the
prediction of rock failure and fracture propagation. In hard rock such as granite, fluid flow
occurs predominately through existing fractures rather than through the rock matrix due to
the rock’s low permeability. Fluid pressure in rock fractures may cause rock fracture
movement, increase fracture aperture or even cause fracture propagation. Fracture
propagation will change the fluid conductivity and create new flow paths. Thus, the
interaction between fracture mechanical response and fluid flow is critical in predicting
rock failure.

Two fundamental approaches have been used to model hydro-mechanical action in


rock medium. The first is an implicit approach employing finite element methods
incorporating Darcy’s Law and fluid flow models through porous media. The second is
an explicit numerical approach where both fluid flow and mechanical response are
simulated using a time step iterative process.

A Numerical Solution Strategy for the Fracturing-Hydraulic Flow Process


An example of the latter approach is a numerical solution strategy for the fracturing-
hydraulic flow process implemented in the computer fracture modeling system, FRACOD,
developed by Shen, B. et. al. (2014). While its details are beyond the scope of this thesis,
the basic solution strategy is as set out below:

Step 1
Fluid flow occurs between fracture domains and fluid leaks into rock matrix. The fluid
flow between the fracture domains is calculated using the Cubic Law (Witherspoon, P.A.
et. al. 1980) in which ‘fluid transmissivity’ is proportional to the cube of aperture and the
pressure gradient. More specifically, the flow rate (Q) between two domains is calculated
as follows:

𝑒𝑒 3 ΔP
𝑄𝑄 = �12𝜇𝜇� • � 𝑙𝑙 � (1.47)

Where:
Q: Flow rate
𝑒𝑒: Fracture Hydraulic Aperture
𝜇𝜇: Fluid Viscosity
ΔP: Fluid Pressure Gradient
𝑙𝑙: Element Length

72
The leakage from the fracture domain into the rock matrix (𝑄𝑄𝐿𝐿𝐿𝐿𝐿𝐿𝐿𝐿 ) is calculated as follows:

𝑘𝑘 𝑃𝑃−𝑃𝑃0
𝑄𝑄𝐿𝐿𝐿𝐿𝐿𝐿𝐿𝐿 = �𝜇𝜇� • � � (1.48)
𝑑𝑑

Where:
𝑄𝑄𝐿𝐿𝐿𝐿𝐿𝐿𝐿𝐿 : Leakage from Fracture Domain into Rock Matrix
𝑘𝑘: Rock Permeability
𝜇𝜇: Fluid Viscosity
𝑃𝑃: Fracture Domain Fluid Pressure
𝑃𝑃0 : Initial Pore Pressure
𝑑𝑑: Effective Leakage Distance

Step 2
Fluid flow causes changes in domain fluid pressure. The new domain pressure due to fluid
flow during a small time duration (Δt) is calculated as follows:

Δ𝑡𝑡 Δ𝑡𝑡
𝑃𝑃(𝑡𝑡 + Δ𝑡𝑡) = 𝑃𝑃0 + 𝐸𝐸𝑤𝑤 • 𝑄𝑄 • � V � − 𝐸𝐸𝑤𝑤 • 𝑄𝑄𝐿𝐿𝐿𝐿𝐿𝐿𝐿𝐿 • � V � (1.49)
Where:
𝑃𝑃: Domain Pressure at time (𝑡𝑡 + Δ𝑡𝑡)
𝑃𝑃0 : Initial Pore Pressure
𝐸𝐸𝑤𝑤 : Fluid Bulk Modulus
Q: Flow rate
𝛥𝛥𝛥𝛥: Time Step
V: Fracture Domain Volume
𝑄𝑄𝐿𝐿𝐿𝐿𝐿𝐿𝐿𝐿 : Leakage from Fracture Domain into Rock Matrix

Step 3
Changes in fluid pressure cause fracture deformation. The fracture deformation is
calculated using Displacement Discontinuity Method (“DDM”) (Shen, B. et. al 2014)
where the new fluid pressures in fracture domains are the input boundary stresses. After
considering the fluid pressure in the fracture domain (elements), the system of equations
for calculating the element displacement discontinuities is given below:

i N ij j N ij j i 
σ s = ∑ Ass Ds + ∑ Asn Dn − (σ s ) 0 
j =1 j =1 
N ij j N  i=1 to N
ij j
(1.50)
σ n = ∑ Ans Ds + ∑ Ann Dn − (σ n ) 0 
i i

j =1 j =1 

73
i i
where σ s and σ n represent the shear and normal stresses of the ith element respectively;
i i
(σ s ) 0 , (σ n ) 0 are the far-field stresses transformed in the crack shear and normal directions.
ij ij j j
Ass , ... , Ann are the influence coefficients, and Ds , Dn represent displacement
discontinuities of jth element which are unknowns in the system of equations.

Step 4
Fracture deformation alters the domain volume, changing the fluid pressure in domains.
The new domain pressure is calculated as follows:

Δ𝑒𝑒•𝑙𝑙
𝑃𝑃′(𝑡𝑡 + Δ𝑡𝑡) = 𝑃𝑃(𝑡𝑡 + Δ𝑡𝑡) − 𝐸𝐸𝑤𝑤 • � V
� (1.51)
Where:
𝑃𝑃′: New Domain Pressure at time (𝑡𝑡 + Δ𝑡𝑡)
𝑃𝑃: Domain Pressure at time (𝑡𝑡 + Δ𝑡𝑡)
𝐸𝐸𝑤𝑤 : Fluid Bulk Modulus
𝑒𝑒: Fracture Hydraulic Aperture
𝛥𝛥𝛥𝛥: Time Step
𝑙𝑙: Element Length
V: Fracture Domain Volume

The new domain fluid pressures are then used to calculate the flow rate between domains
in Step 1. Steps 1-4 are iterated until the desired fluid time is reached and a stable solution
is achieved.

Mathematically Simpler But Requires Significant Computing Time


While the Shen B. et. al. (2014) solution strategy is mathematically simpler than
implicit models of fluid flow through porous media, it does require small time increments
and correspondingly long computing times to reach a stable solution.

Future Developments Likely to Come From Oil Industry’s Hydraulic Fracturing Research
In summary, the rock properties most important to predicting cutting efficiency are
likely to be structural and not physical. The presence of fluids from the water jet represents
an additional layer of complexity. While some physical properties (e.g. compressive
strength, hardness, etc.) are correlated, to a degree, to structural properties they are likely
to serve only as crude indicators of the amenability of a rock mass to cutting by water jet.
A better understanding of the conditions that govern fracture initiation, growth and the
pressurization of such fractures by the water jet fluid is needed to develop enhanced models

74
of the physics of rock cutting by means of water jet. Such knowledge is most likely to
come from ongoing developments in hydraulic fracturing in the oil industry rather than
from the mining industry. The oil industry is spending tens of billions of dollars each year
on hydraulic fracturing of low permeability shale formations (with permeability typically
measured in the nano-darcys) and associated research on elastic fracture mechanics, fluid-
rock interactions and strength anisotropy of low permeability shale formations. The astute
mining engineer will be best served by transferring advances in the oil industry’s enhanced
understanding of the physics of hydraulic fracturing to hydraulic borehole mining.

3.10 Pulsating Jets


Pulsating Jets Allow for Spent Jet to Escape Impact Crater
The idea of interrupted or pulsed jets as a means of improving jet cutting
performance arose from initial studies in the 1960s of the early stages of water droplet
impact (Brook and Summers 1969). Experiments indicated that when a continuous jet of
water is interrupted to produce a succession of short lengths of moving liquid, there is an
increase in penetration. This effect was attributed to the fact that a continuous jet must
overcome the resistance of the escaping water from the impact crater as the jet continues
to penetrate. By contrast, an interrupted flow allows for some escape of the spent jet
backflow thereby clearing access to the target material for the subsequent slug of fluid.

Water Hammer Pressure Plays a More Important Role


Subsequent investigators (Brunton and Rochester 1979) identified a further, more
important factor to explain the superior cutting performance of pulsating water jets, namely
Water Hammer Pressure. When the motion of a liquid segment of a regular coherent jet is
abruptly stopped by a solid body, a very high pressure in the vicinity of the contact area is
created for a very short duration. The initial momentary high pressure results from a “water
hammer” effect, the compressibility of the liquid giving an instantaneous pressure of the
order:

𝑃𝑃𝐻𝐻 = 𝜌𝜌0 • 𝐶𝐶0 • 𝑣𝑣 (1.52)

Where:

𝑃𝑃𝐻𝐻 Water Hammer Pressure


𝜌𝜌0 Density of the incompressible liquid
𝐶𝐶0 Acoustic speed of the liquid
𝑣𝑣 Normal component of the collision velocity

75
After Instantaneous Time Period Hammer Pressure Decays to Stagnation Pressure
Pressure then decays to the ordinary hydrodynamic stagnation pressure given for
incompressible flow as previously presented as Eq. (1.14):

1
P = 2 • ρ • 𝑉𝑉0 2 (1.14)

Where:
P Pressure of Water Jet (N/𝑚𝑚2 )
ρ Mass density of the jet fluid (kg/𝑚𝑚3 )
𝑉𝑉0 Exit velocity of the water jet (m/sec)

Duration of Peak Hammer Pressure


The duration of the peak pressure is estimated to be:

2𝑙𝑙
(1.53)
𝐶𝐶0

Where:
𝑙𝑙 Length of the jet segment
𝐶𝐶0 Acoustic speed of the liquid

Modified Water Hammer Equation to Account for Liquid Incompressibility


The conventional “water hammer” equation must be corrected to allow for the
incompressibility of the liquid which occurs at higher impact velocities.

Hwang and Hammitt (1977) found that use of the following equation,

C v v
C0
= 1 + 2 • �C � − 0.1 • (C )2 (1.54)
0 0

gives less than a 3% error over the range of impact velocity ratios

v
C0
≤3 (1.55)

Beyond this level, “water hammer” pressure reverts to Eq. (1.52), namely:

𝑃𝑃𝐻𝐻 = 𝜌𝜌0 • 𝐶𝐶0 • 𝑣𝑣 (1.52)

Where:
𝑃𝑃𝐻𝐻 Water Hammer Pressure

76
𝜌𝜌0 Density of the incompressible liquid
𝐶𝐶0 Acoustic speed of the liquid
𝑣𝑣 Normal component of the collision velocity

Order of Magnitude Difference between Hammer and Stagnation Pressures


The difference in effective pressure on the target surface developed by a continuous
jet (Eq. 1.14, “Stagnation Pressure”) vs. an interrupted flow (Eq. 1.52, “Water Hammer
Pressure”) is rather striking (pardon the pun). As an example, the Stagnation Pressure
developed by a jet moving at 500 m/sec is about 9% of the pressure developed by the
impact of the head of the jet pulse (Mazurkiewicz 1983). In other words, the level of the
“water hammer” pressure could only be reached by the continuous jet if the pump pressure
is increased about 11 times. This implies a power increase of about 20-30 times after taking
into account the energy losses for a typical system.

Example of Hammer vs. Stagnation Pressures


An expanded comparison of Stagnation vs. Hammer Pressure across a range of
nozzle discharge pressures is shown in Table 4 below. To construct this table the assumed
discharge velocity has be estimated from:

V = 12 • √P (1.56)

Where:
V Nozzle discharge velocity (ft/sec)
P Nozzle Discharge Pressure (psig)

Using Eq. (1.14) and Eq. (1.52), we can construct the following table of calculated
Stagnation and Hammer pressures:

Pulsed Jets: Stagnation vs. Hammer Pressure

Nozzle Discharge Pressure Velocity Stagnation Pressure Hammer Pressure Hammer/Stagnation


(psig) (ft/sec) (psig) (psig) Ratio
350 220 340 15000 44
550 280 530 19000 36
10000 1200 9700 81000 8
30000 2100 29000 140000 5

Table 4 – Hammer vs. Stagnation Pressure for a Pulsed Jet


(Nebeker, 1987)

77
Hammer Pressure can Exceed Stagnation Pressure by 40 Times
As noted from the above table, the Hammer Pressure is significantly greater than
the Stagnation Pressure in each case, varying from 44 to 5 times the Stagnation Pressure.
Incidentally, most rock materials have uniaxial tensile strengths that vary between 500 psig
and 6,000 psig. Compressive strengths, however typically range from 7,000 psig and
70,000 psig. This illustrative exercise suggests the possibility of modifying a conventional
jet of only 10,000 psig discharge pressure into a Pulsed Jet with impact energy close to the
compressive strength of harder rocks such as granite.

Droplet Impact Also Generates High Velocity Lateral Micro-Jets


The augmentation of pressure does not end with the above noted “hammer
pressure”, nor is it an even distribution of pressure over the target surface. As the curved
droplet surface hits the target material surface, the initial contact in the center sends out a
compression wave through the drop. (See Figure 34 below). This augments the impact
velocity and hastens the downward movement of sequential rings of contact around the
initial point. At the same time these rings have a shorter distance to travel to reach the
surface than does the water at the initial point of contact which is now trying to escape
from the slug of fluid flowing down behind it. This confinement increases pressure on the
contact surface and the fluid at the surface. Pressure augmentation continues until the point
where the curvature of the drop raises the travel distance of the next ring of fluid above
that needed to allow the confined fluid to escape, which it does. The resulting lateral micro-
jet, presuming no significant target penetration, travels at a much higher velocity than the
original impact. Fyall (1967) measured radial flow velocities some three times the initial
300 m/sec impact velocity of the droplets.

78
Figure 34 – Fluid Droplet Collapsing Against a Wall
(Benjamin & Ellis 1966)

Water Hammer is the Single Most Important Cutting Force


In sum, Water Hammer Pressure is the more important of the two above mentioned
phenomena and can exceed the stagnation pressure of a continuous water jet by an order
of magnitude. As such, a pulsed jet stream can cut rocks which require cutting pressures
much greater than the available hydraulic pressure. While the high lateral velocity micro-
jets are an important cutting force, the water film over the target tends to mitigate somewhat
the full scouring effect as compared to the water hammer effect.

3.11 Cavitating Jets

Pressure Augmentation Can Also be Achieved by Cavitation


The benefits of pressure augmentation described in the preceding Section can also
be achieved in another way. This is by taking advantage of the power of “cavitation bubble
collapse”. This phenomenon has been known for many years for its destructive impact on
pumps, propellers, fluid delivery lines, etc. and was first examined theoretically by
Benjamin and Ellis (1966). Subsequent work in the 1980s by researchers at Hydronautics
Inc. (e.g. Chahine, et. al. 1982) highlighted the potential application of this physical
phenomenon as a promising cutting force.

Cavitation Occurs When Water is Stretched


Cavitation occurs when, through one process or another, water is stretched. Because
water isn’t very strong in tension, too much of a pull will cause it to form small vacuum
bubbles within the flow (cavities), as the water is ripped apart. These cavities in the flow
aren’t stable and rapidly collapse as soon as the surrounding fluid reaches any significant
pressure. The bubbles themselves are individually tiny, but their destructive power can be
quite dramatic.

79
Collapsing Bubbles Generate Destructive Forces
The destructive force of cavitation occurs when the bubble collapses as it does so
asymmetrically. It folds over (term is “involute”) on itself and a tiny jet known as a Munroe
jet is formed by the convergence of the collapsing walls. (See Figure 35 below). The size
of the bubbles and the micro-jets thus created are very small. Typical bubble diameters are
on the order of 10 to 25 μm with the resulting micro-jets being about one-tenth of this size
(Summers 1995).

Figure 35. Schematic Showing the Development of a Micro-jet within the Collapsing Wavity
(Summers: Bit Tooth Energy Blog, 2016)

While Small, Cavitation Bubble Collapse is a Violent Process


Cavitation bubble collapse is a violent process that results from the inherent
instability of the bubbles due to their non-spherical form and intrinsic pressure differentials.
(See Figure 36 below). When the bubbles collapse, they generate highly localized, large-
amplitude shock waves in the fluid at the point of collapse. When this collapse occurs close
to a solid surface, these intense disturbances generate highly localized and transient surface
stresses. (See Figure 37 below). Repetition of this loading due to repeated collapses causes
local surface fatigue failure and the subsequent detachment or flaking off of pieces of
material. This is the generally accepted explanation of cavitation damage.

80
Figure 36. High Speed Photos Showing Stages of Bubble Collapse
(Summers: Bit Tooth Energy Blog, 2016)

81
Figure 37 - Series of photographs of a cavitation bubble collapsing near a wall along with the characteristic
wall pressure trace. The time corresponding to each photograph is marked by a number on the trace. From
Shima, Takayama, Tomita, and Ohsawa (1983)

Collapsing Pressure is on the Order of 1 Million psi


The typical collapsing pressure jet from a cavity collapse is on the order of 1 million
psi (El Saie 1977, Summers 2016). This reportedly occurs with relatively little control by
the surrounding fluid or originating jet and implies that it is quite possible to use higher
volume jet streams and still achieve quite destructive results. As a further example of the
destructive potential of cavitation, Summers (2016) reports that a jet pressure of 1,600 psi
was able to drill through blocks of granite with a typical compressive strength of 20,000
psi.

Cavitation Induction in a Nozzle


Cavitation is created by developing shear or tensile forces in the fluid stream. In
many nozzle designs, cavitation bubbles are formed in a high speed jet by placing a flat
ended probe in the path of the nozzle, so arranged that the flat end is situated beyond the
acceleration section of the nozzle (see Figure 38 below). In this way, as the jet reaches
maximum speed and moves from the discharge outlet of the nozzle, it pulls a vacuum in
the fluid directly over the tip of the probe generating small cavitating bubbles that are then
drawn into the jet flow. The bubbles are carried by the main jet to the target where they
collapse on the surface of the target material, creating small craters. Under sustained
action, the craters will gradually coalesce, creating cracks and ultimately tear away the
surface of the target material.

Figure 38. Design of a Nozzle with Center Body to Induce Cavitation


(Summers 1995)

82
Cavitation Holds Great Promise
The fact that cavitation allows rock to be cut at a fluid pressure significantly below
the stagnation pressure of a continuous high pressure jet offers great promise. This is one
of the most promising areas of future research, particularly at significantly higher flow
rates than achieved to date and over a wider spectrum of minerals (since the destructive
pressures generated by cavitation exceed those necessary to disintegrate even the strongest
of natural minerals).

3.12 Extending the Cutting Range – Introduction of Abrasives

Extending the Cutting Range for Harder Rocks


The cutting effectiveness of a water jet is limited, in part, by the hardness of the
target material. In the case of kimberlite ores, Merlin’s (2012) lab testing suggests that ore
with a specific gravity of more than 2.3 g/cc will be too hard to cut by water jet and
therefore not amenable to HBHM. This effectively limits HBHM to the weathered sections
of a kimberlite ore body. This restricts the quantity of ore that can be mined, adversely
impacting project economics.

Introduction of Abrasives Permits the Cutting of Harder Rocks


To extend the range of kimberlite ores amenable to HBHM, we consider the
introduction of abrasives into the water jet stream. This technique has been successfully
applied in other water jet applications such as industrial cleaning, metal and glass cutting
and hydro-demolition of concrete surfaces and reinforced concrete. To the best of our
knowledge it has not yet been tested on the mining of harder ore types. While a full
technical treatment of the topic is beyond the scope of this thesis, we undertake in this
section an introduction of the underlining physics and operating considerations of an
abrasive water jet system.

Injection of Abrasive Particles into the Water Jet


An abrasive water jet is formed by accelerating small solid particles (garnet, silica
sand, aluminum oxide, etc.) into a mixing chamber where the particles are entrained in the
water jet. In a typical design, a high pressure stream of water enters a mixing chamber
through the upper nozzle, passes through the chamber, creating a vacuum that pulls
abrasive into the chamber, and mixes with that abrasive before exiting the nozzle toward
the target rock. (See Figure 39).

83
Figure 39 – Basic Abrasive Water Jet Design
(Source: www.bittooth.blogspot.com Dec. 2013)

Rock is Considered a Brittle Material


As a general rule, rock is considered to be a brittle material, although there are some
rock types within this classification (for example some shales) which could more
appropriately described as ductile, both in overall behavior and their response to particle
impact. The failure of brittle materials is quite different to that of ductile materials and
occurs by a process of crack creation and extension. This has been studied in detail for
abrasive water jets, particularly for glass cutting which is a hard, brittle material (Summers
1995). Glass has the added advantage of allowing visual observation.

Mechanism for Rock Failure


As originally described by Hertz (1882), crack formation is the result of dynamic
loading of the surface of the target material by the impacting abrasive particle. The crack
formation in the case of a spherical particle follows a distinct path. Initially the surface of
the rock impacted by the particle is compressed and must stretch to maintain contact with
the crater being created by the impacting particle. As the dynamic load increases, and the
size of the crater grows, radial cracks are generated in the stretched zone of material around
the perimeter of contact. Continued dynamic loading also leads to the creation of lateral
cracks, running almost parallel to the surface, but within the bounds of the radial cracks.
As the force on the surface increases, lateral cracks continue to extend and begin to join
with other cracks, ultimately joining together to isolate a block of rock leading to complete
failure and material removal.

84
Principal Stresses Define Direction of Crack Growth and Propagation
The crack path is defined by principal stress trajectories. The initial cone crack is
orthogonal to the highest principal stress in the stress field prior to cracking. The crack
then propagates along the planes defined by small principal stresses. Ideally, the crack
tries to propagate in a straight line. However, oblique stresses are generated on the crack
when the direction of the principal stresses changes. These stresses force the crack to
correspond to the principal stress directions and propagate in an unstable manner.

Hertzian Theory for Elastic Contact


The modeling of crack formation and propagation was first postulated by Hertz
(1882) and accordingly is referred to as the Hertzian theory for elastic contact. As defined
by Hertz, tensile stresses generated by an impinging spherical particle are maximum at the
surface of the edge of contact, namely:

(1−2•VM )•FC
𝜎𝜎𝑇𝑇 = (1.57)
2•𝜋𝜋•𝑟𝑟 2

Where:
𝜎𝜎𝑇𝑇 : Maximum Tensile Stress
VM : Volume Removal
FC : Contact Force
𝑟𝑟: Contact Radius

Maximum Tensile Strength is Related to Particle Velocity


Equation (1.57) gives rise to the following general relationship between maximum
tensile stress and particle velocity (Sheldon and Finnie 1966):

2/5
𝜎𝜎𝑇𝑇 ∝ 𝜈𝜈𝑝𝑝 (1.58)

Where:
𝜎𝜎𝑇𝑇 : Maximum Tensile Stress
𝜈𝜈𝑝𝑝 : Particle Velocity

As Equation (1.58) clearly shows, increasing particle velocity results in higher tensile
stress.

Particle Velocity Influences Stress Which Then Impacts Crack Length and Hole Depth
Higher particle velocity results in higher tensile stress which, in turn, leads to
greater crack length and hole depth. The direct relationship between impact velocity and
stress-induced radial crack length and hole depth is empirically shown in Figures 40 and

85
41 below. As impact velocity increases, so does tensile strength and the resulting crack
length and hole depth.

Figure 40 - Effect of Impact Velocity on Crack Length


(Evans 1979)

Figure 41 - Berea Sandstone Removed by the Impact of Steel Balls of Varying Size
(Ripkin, Wetzel 1972)

86
Erosion is Function of Impact Angle, Particle Size, Velocity, Existing Cracks, etc.
Damage by the incoming abrasive particle is most intense at high angles of impact,
with perpendicular impacts rendering the maximum damage. Theoretical work by Sheldon
and Finnie (1966) and Evans (1979) assumes that erosion occurs as the result of Hertzian
contact stresses during impact. These stresses cause cracks to grow from pre-existing flaws
in the target rock surface. The load at which crack propagation occurs is related to the
distribution of surface flaws which in turn are characterized by Weibull distribution
parameters describing existing cracks in the target material. The theoretical and
experimental work by Sheldon and Finnie (1966), Evans (1979) and others has resulted in
the following generalized predictive model of the volume of eroded material removed from
an impacting particle:

C
VM = C1 • 𝑟𝑟 2 • 𝜈𝜈𝑝𝑝 2 (1.59)

Where:
VM : Volume Removal
C1 : Constant depending on physical characteristics of impacting particle (density) and
target material (Weibull factors, modulus of elasticity)

𝑟𝑟: Particle Radius


𝜈𝜈𝑝𝑝 : Particle Velocity
C2 : Constant related to crack length and depth

Theoretical values predicted by Equation (1.59) have been compared with experimental
data and been found to be in reasonable agreement (Evans 1979). The basic premise is that
the volume of material that will be eroded is defined by the bounds of the damaged zone,
notably its diameter and depth.

Crack Distribution is Important but Difficult to Model


In the case of rocks, the amount of material removed as a result of impact will
depend significantly on the density and distribution of the cracks which exist prior to the
impact. This adds immensely to the complexity of modeling erosion resulting from particle
impact. While the inclusion of Weibull distribution constants for crack size and density
distribution implicit in Equation (1.59) is meant to approximate this aspect of the target
rock, it is only a crude approximation. Further refinements and comprehensive predictive
models have been proposed but all suffer limitations. Additional research is warranted to
further the understanding of brittle erosion processes.

87
Kinetic Energy Consists of Two Components: Particles & Water
The kinetic energy of a hydro-abrasive water jet is defined as:

𝑁𝑁 𝑚𝑚𝑤𝑤
E = ∑𝑖𝑖 𝑝𝑝 E𝑝𝑝𝑝𝑝 + • 𝜈𝜈𝑝𝑝 2 (1.60)
2

Where:
E: Kinetic Energy of Hydro-Abrasive Water Jet
E𝑝𝑝𝑝𝑝 : Kinetic Energy of Abrasive Particle i

𝑁𝑁𝑝𝑝 : Number of Abrasive Particles


𝑚𝑚𝑤𝑤 : Water Mass Flow Rate
𝜈𝜈𝑝𝑝 : Particle Velocity

Equation (1.60) implicitly assumes that the abrasive particles and water phase in the hydro-
abrasive jet have equal velocities. With this assumption, the left term of Equation (1.60)
is the energy delivered by the abrasive particle to the target rock surface. This portion is
typically 10% of the total kinetic energy of a hydro-abrasive water jet. The remaining 90%
is carried by the water phase of the jet (Momber 2001).

Significant Energy is Lost Accelerating the Abrasive Particles


At the levels of abrasive feed normally employed, there is significant energy loss
from the water jet in accelerating the abrasive particles. The greater the number of abrasive
particles, the slower the water jet for a given jet pressure. Since the damage caused by
individual particles is related to impact velocity, the slower the particles are traveling, on
a unit volume basis, the less target material removed.

Economic Trade-Off Between Increasing Particle Energy vs. Erosion Gains


By the same logic, increasing jet pressure increases energy available to accelerate
the particles, and the erosion therefore increases. See Figure 42 Below. This directly
increases the amount of material removed by each abrasive particle but at the same time
increases the overall energy required to remove a unit volume of material. This then
represents a key trade-off in practice and needs to be evaluated for each specific
application; namely the effective cost of mining (US$/tonne) for varying pressure rates.

88
Figure 42 - Effect of Increase in Jet Pressure on Cut Depth
(Hashish 1991)

Increasing the Abrasive Feed Rate Improves Cutting


At low levels of abrasive feed, Hashish (1991) has shown that increasing the
amount of abrasive in the feed increases cutting performance. See Figure 43 below.

Figure 43 - Effect of Increase in AFR on Depth of Cut in Mild Steel

(Hashish 1991)

89
However, There are Limits…
However, as the abrasive flow continues to increase, the cutting performance
reaches a limit and then declines (Hashish 1991). See Figure 44 below.

Figure 44 - The Effect of Higher AFR on Cutting Depth at 3 Jet Pressures on a Mild Steel Target
(Hashish 1991)

Choice of Abrasive Particle Type Critical to Performance


Aside from the normal operating parameters affecting the performance of a water
jet (e.g. nozzle size, stand-off distance, traverse rate, pressure, etc.), the choice of abrasive
particle is critical to ensuring optimal performance of an abrasive water jet system. This
relates not only to particle size, density and hardness but also to particle shape. A spherical
shape is preferred in the case of cutting rock as it best concentrates impact energy to
maximize the generation of cracks in the rock. While there is considerable research and
experience with abrasive selection for industrial cleaning, hydro-demolition and various
machining applications, little or no research exists for HBHM and rock cutting. This
represents a fruitful area of future research.

Cost is an Important Factor


The cost of various abrasive options is a further consideration. As an example,
while small, industrial diamond grit may yield the best results their cost is likely to be
prohibitive. A related consideration is the impact of any chosen abrasive on performance
of the air lift system and potential downstream complications in separating out the chosen
abrasive from the economic ore of interest. Again, both of these considerations would

90
benefit from further investigation as there seems to be little or nothing at present in the
research literature.

Abrasives Offer Great Potential but Many Open Questions


In conclusion, the introduction of abrasives into the water jet stream offers the
potential to greatly expand the range of ores that can be cut by HBHM, particularly for
harder ore types like kimberlites. However, there is little or no historic experience with
the use of hydro-abrasive jets in HBHM applications. This is compounded by the many
challenges in developing robust predictive models, especially for the central question of
cutting rate. For the foregoing reasons, it is likely that in the near term progress will only
come from field experience and empirically derived models of abrasive water jet
performance.

3.13 Airlift

Airlift – Simple & Efficient But with Limitations


Use of airlift technology within an HBHM system allows for the efficient lifting of
the mined ore-water slurry mix from the cutting face to the surface. As part of a
comprehensive HBHM system, airlift has the added advantage of allowing the full pressure
pumping capacity to be allocated to rock cutting and not partially diverted for lifting
purposes. Airlift pumps have a number of advantages over mechanical pumps, the
principal ones being low cost, simplicity, efficiency and reliability. Airlifts can be
particularly useful (almost essential) where corrosive, abrasive or radioactive fluids must
be transported. Like all technology solutions, however, airlift pumps have their limitations
including limited suction and lift capacity, weak flow control, unstable flow rates and a
vulnerability to clogging.

Rising Air Lowers Fluid Density Creating a Suction Effect


The concept of an airlift pump was established in the late 1800’s. The basic
principle involves the release of compressed air at the bottom end of a vertical pipe within
an enclosed tube or pipe. The air, once released in a conduit, begins to rise and expand
thereby creating a suction effect at the bottom of the pipe by lowering the density of the
fluid mixture. As the air bubbles rise and expand on their upward return journey they
continue to lower the specific gravity of the surrounding fluid mixture providing a
sustaining pull of the entire fluid mixture to the surface. An illustration of a simple airlift
pump is shown in Figure 45 below.

91
Figure 45 – Illustration of Simple Airlift Pump
(Source: www.plantservices.com)

Bernoulli’s Equation Explains Underlying Physics of Airlift Flow


The physics of airlift flow through a pipeline is governed by Bernoulli’s equation
which is a statement of conservation of energy applied to fluid flow through a pipeline. At
each point along the pipeline the total energy of the fluid is computed by taking into
consideration the fluid energy due to pressure, velocity and elevation combined with any
energy input, energy output and energy losses. The total energy of the fluid contained in
the pipeline at any point is a constant. This is the well known principle of conservation of
energy.

Figure 46 – Total Energy of Fluid in Pipe Flow


(Source: www.teachengineering.org)

92
Derivation of Bernoulli’s Equation
Consider a fluid flow through a pipeline from Point 1 to Point 2 as shown in Figure
46. The elevation at Point 1 is ℎ1 and the elevation at Point 2 is ℎ2 above a common
reference point, e.g. ground level (i.e. ℎ = 0). The pressure at Point 1 is 𝑃𝑃1 and at Point 2
is 𝑃𝑃2 . It is assumed that the pipe diameters at Points 1 and 2 are different, and hence the
flow velocity at 1 and 2 will be represented by 𝑣𝑣1 and 𝑣𝑣2 , respectively. A particle of the
fluid of unit weight at Point 1 in the pipeline possesses a total energy E which consists of
three components:

Potential Energy = ℎ1
𝑃𝑃1
Pressure Energy = ρ
𝑣𝑣1 2
Kinetic Energy = ,
2𝑔𝑔

where ρ is the density of the fluid (kg/𝑚𝑚3 ) and 𝑔𝑔 is the acceleration due to gravity (m/𝑠𝑠 2 ).

Therefore the total energy at Point 1 is:

𝑃𝑃1 𝑣𝑣1 2
𝐸𝐸 = ℎ1 + ρ
+ 2𝑔𝑔
(1.61)

Since each term in Eq. (1.61) has dimensions of length we refer to the total energy at Point
1 as 𝐻𝐻1 in meters of fluid head. Therefore, rewriting the total energy in meters of fluid
head at Point 1, yields:

𝑃𝑃1 𝑣𝑣1 2
𝐻𝐻1 = ℎ1 + + (1.62)
ρ 2𝑔𝑔

Similarly, the same unit weight of fluid at Point 2 has a total energy per unit weight of 𝐻𝐻2
given by:

𝑃𝑃2 𝑣𝑣2 2
𝐻𝐻2 = ℎ2 + + (1.63)
ρ 2𝑔𝑔

By the principle of conservation of energy:

𝐻𝐻1 = 𝐻𝐻2 (1.64)

93
Therefore,

𝑃𝑃1 𝑣𝑣1 2 𝑃𝑃2 𝑣𝑣2 2


ℎ1 + + = ℎ2 + + (1.65)
ρ 2𝑔𝑔 ρ 2𝑔𝑔

In Eq. (1.65), referred to as Bernoulli’s equation, we have not considered any energy
added to the fluid, energy taken out of the fluid, or energy losses due to friction.
Therefore, modifying Eq. (1.65) to take into account the addition of energy (such as
from a pump at Point 1) and accounting for frictional head losses, ℎ𝑓𝑓 , we get the more
common form of Bernoulli’s equation as follows:

𝑃𝑃1 𝑣𝑣1 2 𝑃𝑃2 𝑣𝑣2 2


ℎ1 + + + 𝐻𝐻𝑝𝑝 = ℎ2 + + +ℎ𝑓𝑓 , (1.66)
ρ 2𝑔𝑔 ρ 2𝑔𝑔

where 𝐻𝐻𝑝𝑝 is the equivalent pressure added to the fluid by pump at Point 1 and ℎ𝑓𝑓 represents
the total frictional pressure losses between Points 1 and 2.

Pipeline frictional pressure losses from Points 1 to 2, ℎ𝑓𝑓 , can be calculated using the Darcy-
Weisbach equation:

L v2
ℎ𝑓𝑓 = 𝑓𝑓 • •
D 2𝑔𝑔
(1.67)

Where:

ℎ𝑓𝑓 Frictional pressure loss (N/𝑚𝑚2 )


𝑓𝑓 Darcy friction factor, dimensionless
L Pipe Length (m)
D Inside Pipe Diameter (m)
v avg. flow velocity (m/s)
𝑔𝑔 acceleration due to gravity (m/𝑠𝑠 2 )

Substituting Eq. (1.67) into Eq. (1.66) yields:

𝑃𝑃1 𝑣𝑣1 2 𝑃𝑃2 𝑣𝑣2 2 L v2


ℎ1 + + + 𝐻𝐻𝑝𝑝 = ℎ2 + + + (𝑓𝑓 • • ) (1.68)
ρ 2𝑔𝑔 ρ 2𝑔𝑔 D 2𝑔𝑔

Eq. (1.68) essentially says that the total energy per weight of fluid at Point 1 is equivalent
to the total energy per weight of fluid at the downstream Point 2 plus the energy losses
between Points 1 and 2.

94
Niklin’s (1963) Model Provides a Useful Predictor of Airlift Performance
Although the principles of an airlift pump are simple, the theoretical study of its
performance is complex and dependent on numerous factors including, but not limited to,
fluid-solid mix type, flow regime, fluid density, pipe diameter, pipe length, air flow rate
and pressure, fluid flow rate and pressure, submergence ratio, boundary layer
considerations, etc. The first comprehensive theoretical model of the air lift pump was
presented by Nicklin (1963). He studied the effects of different parameters including
diameter, length, pressure at the top of the riser tube, submergence ratio, and water
volumetric flow rate on airlift pump efficiency. Nicklin’s analysis concluded that by
neglecting the entrance effects and assuming slug flow in the riser tube, the performance
of the airlift pump can be predicted based on a two-phase slug flow regime. Moreover, he
established a definition of efficiency as the work done in lifting the fluid, divided by the
work done by the air as it expands isothermally, i.e.

ρ•g•𝑄𝑄𝐿𝐿 •(L−𝐻𝐻𝑠𝑠 )
ɳ= 𝑃𝑃 (1.69)
𝑃𝑃𝑎𝑎 •𝑄𝑄𝑎𝑎 •Ln• 𝑖𝑖𝑖𝑖
𝑃𝑃𝑎𝑎

Where:

ɳ Airlift Pump Efficiency (%)


ρ Fluid Density (kg/𝑚𝑚3 )
g Gravitational Acceleration (m/𝑠𝑠 2 )
𝑄𝑄𝐿𝐿 Discharge Volume Flow Rate (𝑚𝑚3 /s)
L Pipe Length (m)
𝐻𝐻𝑠𝑠 Static Depth of Water (m)
𝑃𝑃𝑎𝑎 Atmospheric Pressure (N/𝑚𝑚2 )
𝑄𝑄𝑎𝑎 Air Volumetric Flow Rate (𝑚𝑚3 /s)
𝑃𝑃𝑖𝑖𝑖𝑖 Air Injection Pressure (N/𝑚𝑚2 )

Other Researchers Have Extended Nicklin’s Seminal Work


The theoretical framework developed by Nicklin (1963) was extended by
Reinemann et. al. (1986) by incorporating the effect of surface tension on bubble velocity.
A further analytical study of airlift pump performance was presented by Stenning and
Martin (1968). They employed continuity and momentum equations assuming one
dimensional flow in the pump riser, and used the results of two phase flow research to solve
the governing equations. Their work found that one-dimensional flow theory provides an
improved model for predicting the performance of airlift pumps. Kassab et. al (2008)
developed a more comprehensive predictive model of airlift performance when operating
in a two phase flow regime.

95
Varying Flow Regimes Add Complexity
Functioning of an airlift is not as simple as suggested above. The velocity and
geometry of the components of the three-phase flow (gas, liquid and solids) influence the
flow regime which is occurring within the riser pipe. Considering for the moment only the
case of two-phase flow (i.e. gas and liquid), there are five distinct flow regimes in a vertical
pipe (See Figure 47 Below). A brief description of each follows:

Disbursed Bubbly Flow – numerous bubbles as the gas is dispersed in the form of discrete
bubbles in the continuous liquid phase. The bubbles may vary widely in size and shape
but they are typically spherical and much smaller than the diameter of the pipe.

Slug Flow – with increasing gas void fraction, the proximity of the bubbles is very close
such that the bubbles collide and coalesce to form larger bubbles, which are smaller in
dimension than the pipe diameter. These bubbles have a characteristic shape similar to a
bullet with a hemispherical nose and a blunt tail end.

Churn Flow – increasing the velocity of the flow, the structure of the flow becomes
unstable with the fluid traveling up and down in an oscillatory fashion but with a net
upward flow. This flow pattern is an intermediate regime between slug flow and annular
flow.

Annular Flow – once the inter-facial shear of the high velocity gas on the liquid film
becomes dominant over gravity, the liquid is expelled from the center of the pipe and flows
as a thin layer on the wall while the gas flows as a continuous phase up the center of the
pipe.

Annular Flow with Droplets – at very high gas flow rates, the annular film is thinned by
the shear of the gas core on the interface until it becomes unstable and is destroyed, such
that all the liquid is entrained as droplets in the continuous gas phase, inverse to the bubbly
flow regime.

96
Figure 47 –Flow Regimes
(Source: www.drbratland.com)

Varying Flow Regimes Exist Within Riser Pipe


Varying flow regimes exist at different points in the riser pipe further complicating
modeling of the system. At the point where compressed air is injected in the pipe, a bubbly
flow exists as the large pressure results in very small air bubble diameters. As they progress
upward, the size of the bubbles increases with decreasing pressure causing slug and churn
flow. Further up, when the velocity of the air has increased, annular flow or annular flow
with droplets may occur. Each flow regime has a different lifting efficiency rendering the
modeling of the system non-trivial.

Flow Regime Maps Help But With Notable Limitations


To model flow regime transitions within airlift systems, many researchers have
developed flow regime maps based on theoretical and experimental results. An example
of such a map produced by Taitel et. al. (1980) is shown in Figure 48 below. While helpful,
such maps are limited to the specific design parameters of airlift system under study (i.e.
specific submergence ratio, pipe diameter, air pressure regime, etc.).

97
Figure 48 –Example Flow Regime Map
(Taitel et. al, 1980)

Three-Phase Systems: A Giant Leap in Complexity


While a three-phase system model (i.e. gas-liquid-solids) better represents the
vertical transport of the ore-water slurry produced by the borehole water jet, it comes with
significant added complexity. Giot (1982) proposed a model for a three-phase airlift pump
used in the pumping of metallic nodules from the sea bed floor. Hu et. al. (2012)
established a theoretical model based on momentum theory by considering the mixture
flow governing equations for airlift and obtaining the relationship among the volumetric
fluxes of air, water and solid phases. A comparison of the accuracy of the Hu et. al. (2012)
model with experimental data found that the proposed model was more accurate for the
modeling of air-water than air-water-solid. Pougatch and Salcudean (2008) simulated deep
sea air lift with a mathematical model of the three-phase flow in a vertical pipe and studied
the influence of pipe diameter on airlift efficiency. Numerous small scale experimental
studies of three phase airlift systems have been undertaken by Fujimoto (2004), Kato et.
al. (1975), Weber Dedegil (1976), Yoshinaga and Sato (1996), Kassab et. al. (2007) and
others.

98
A Conceptual Three-Phase Airlift Model
A conceptual configuration of a three-phase airlift pump having uniform cross-
sectional area is presented by Kassab et. al. (2007) and illustrated in Figure 49 below. This
conceptual airlift pump has a pressure distribution, P, in the flow direction, z. The pump
consists of two parts: a suction pipe in which a two-phase water-solid mixture flows and a
riser pipe in which a three-phase mixture of air-water-solids flows. The symbols E, I and
O denote the cross sections of the suction pipe inlet, the air injector and the riser outlet,
respectively.

Figure 49 –Conceptual Airlift Pump and Axial Pressure Distribution


(Kassab et. al, 2007)

Momentum Balance Method Used Most Often for Three-Phase System Modeling
Momentum balance is the method most often used in theoretical studies of three-
phase systems. It uses the loss in density within the vertical pipe and the friction force to
calculate the difference between the mass flux times velocity (i.e. momentum) at the intake
valve and the discharge valve.

99
The momentum equation is applied to a control volume bounded by the pipe wall and the
cross sections, E and O. Assuming the solid particles are of the same size and density, the
momentum equation may be written as:

𝐼𝐼
𝐴𝐴�𝐽𝐽𝐿𝐿 𝜌𝜌𝐿𝐿 𝜇𝜇𝐿𝐿,𝐸𝐸 + 𝐽𝐽𝑆𝑆 𝜌𝜌𝑆𝑆 𝜇𝜇𝑆𝑆,𝐸𝐸 � − 𝐴𝐴�𝐽𝐽𝐺𝐺,𝑂𝑂 𝜌𝜌𝐺𝐺,𝑂𝑂 𝜇𝜇𝐺𝐺,𝑂𝑂 + 𝐽𝐽𝐿𝐿 𝜌𝜌𝐿𝐿 𝜇𝜇𝐿𝐿,,𝑂𝑂 + 𝐽𝐽𝑆𝑆 𝜌𝜌𝑆𝑆 𝜇𝜇𝑆𝑆,𝑂𝑂 � − 𝜋𝜋𝜋𝜋 ∫𝐸𝐸 𝜏𝜏𝐿𝐿𝐿𝐿 𝑑𝑑𝑑𝑑 −
𝑂𝑂 𝐼𝐼 𝑂𝑂
𝜋𝜋𝜋𝜋 ∫𝐼𝐼 𝜏𝜏3 − 𝐴𝐴 ∫𝐸𝐸 �𝜌𝜌𝐿𝐿 𝜀𝜀𝐿𝐿,𝐿𝐿𝐿𝐿+ 𝜌𝜌𝑆𝑆 𝜀𝜀𝑆𝑆,𝐿𝐿𝐿𝐿+ � 𝑔𝑔𝑔𝑔𝑔𝑔 − 𝐴𝐴 ∫𝐼𝐼 �𝜌𝜌𝐺𝐺 𝜀𝜀𝐺𝐺,+ 𝜌𝜌𝐿𝐿 𝜀𝜀𝐿𝐿,3,+ 𝜌𝜌𝑆𝑆 𝜀𝜀𝑆𝑆,3, � 𝑔𝑔𝑔𝑔𝑔𝑔 +
𝑂𝑂
𝐴𝐴 ∫𝐼𝐼 �𝜌𝜌𝐿𝐿,𝑔𝑔 (𝐿𝐿2 + 𝐿𝐿3 )� = 0 (1.70)

where J is the volumetric flux, 𝜇𝜇 the velocity, 𝜌𝜌 the density, 𝜏𝜏 the shear stress, and 𝑔𝑔 is the
gravitational acceleration. The subscripts, G, L, S, LS and 3 represent air, water, solid, two
phase water-solid mixture and three-phase air-water-solid mixture, respectively. The
subscripts E, I and O represent the cross sections of the inlet, air injector and outlet,
respectively.

While somewhat daunting, Eq. (1.70) is more easily understood by parsing each of
its components. The first and second terms of Eq. (1.70) denote the momentum which
enters through E and leaves through O. The third and fourth terms denote the frictional
pressure loss in the two phase water-solid flow and in the three phase air-water-solid flow,
respectively. The fifth and sixth terms denote the weight of the two-phase water-solid
mixture and that of the three-phase mixture, respectively. The seventh and final term
denotes the pressure force of the surrounding water acting on E. The pressure at O is
assumed to be equal to atmospheric pressure. To satisfy the basic principle of conservation
of momentum, the sum of the seven terms must equal zero.

Various simplifying assumptions and an iterative numerical solution strategy are employed
to find solutions for Eq. (1.70). These are beyond the scope of this thesis.

Key Parameters for Optimal Airlift Performance


Kassab’s et. al. (2007) three-phase airlift model is in reasonable agreement with
experimental results obtained over a range of airlift operating parameters and
characteristics. Accordingly, his work provides a reasonable predictive model for
performance of a typical HBHM airlift system, e.g., in the case of this thesis the mining of
weathered kimberlite sections yielding a slurry density of 20-25% solids at flow rates of
1,000-1,500 gpm.
Conclusions derived from the Kassab et. al. (2007) three-phase model are as follows:
• As the submergence ratio increases, the maximum efficiency of the airlift pump
increases assuming a constant airflow rate.
• For the same submergence ratio, varying the length of the riser pipe affects the
airlift pump performance.

100
• The mass flow rate of the solid particles increases with the decrease of the particle
size.
• The performance of the airlift pump, lifting water and sold particles, depends on
the flow pattern in which the pump operates.
• Maximum airlift fluid rate is achieved in slug or slug-churn flow regimes.
• The pressure of compressed air limits the discharge rate of the airlift pump.
• As the diameter of the riser pipe increases, so does the airlift efficiency.
• The three-phase model can be used to predict two-phase system performance by
setting the solid mass flow rate in the model to zero.

Kassab’s Model Can Be Used To Design an Optimal HBHM Airlift System


The three-phase model proposed by Kassab et. al. (2007) is a reasonable predictor
of airlift pump performance and can be used to optimize design parameters of an HBHM
airlift system for differing flow patterns. Results from HBHM pilot testing can then be
used to adjust operating parameters (e.g. air and fluid pressure and flow rates) and system
components (e.g. riser pipe diameter, nozzle size, etc.) to optimize overall system
performance and ensure a smooth and sustained flow of slurry to the surface.

3.14 System Losses in Delivering the Energy

System Pressure Losses Arise From Two Sources


As discussed above, there are two major sources of pressure loss in an HBHM
system. The first source of pressure loss is from within the HBHM delivery system due to
line losses and system component inefficiencies. The second source is the loss in energy
after the jet has discharged from the nozzle. In this section we discuss in turn both of these
sources of energy losses.

Line Losses Are Significant So Care Must Be Taken in System Design


The first source of energy losses come from within the HBHM system as the water
travels from the pump to the nozzle. The energy losses occur through several mechanisms.
Where the flow is directed through sharp turns, energy is consumed to change the direction
of water flow. Where the flow line undergoes a sharp contraction or expansion, additional
energy is lost in the turbulence patterns which are created. The greatest pressure loss will,
however, come from line losses from the inner pipe and hose which are too small for the
volume of water flow.

Valves and Fittings Are Big Culprits


Values for pressure losses which are caused by the water flowing through valves
and fittings should be provided by the manufacturer of the part and should be correlated to
the flow which is passing through the fitting. This information is typically given as a

101
resistance equivalent to a length of pipe of a given inner diameter. This permits the mining
engineer to calculate the pressure loss for different volume flow through the part.

Calculating Pressure Losses in Delivery Line


In calculating the pressure loss in driving the water through the line there are several
methods than can be used. Labus (1989) has suggested the following equation can be used
to determine the pressure loss in the delivery line of most commercial operations where
high pressure water is used:

𝑏𝑏𝑏𝑏𝑏𝑏
𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃 𝐷𝐷𝐷𝐷𝐷𝐷𝐷𝐷 ( 𝑚𝑚 ) = (0.597 • 𝑄𝑄 2 )/( 100 • 𝐷𝐷5 • 𝑅𝑅𝑅𝑅𝑅𝑅 0.25 ) (1.71)

Where:

Q Flow Rate of Water Jet (liters/min)


𝐷𝐷 Pipe diameter (cm)
𝑅𝑅𝑅𝑅𝑅𝑅 Reynolds number given by 1116.5•(Q/D)

Line Losses – A Real Example


Consider a 1.25 cm internal diameter hose carrying a flow of 50 lpm. The Reynolds
Number for such a flow would be (1116.5 X 50/1.25) or 44,660. Employing Equation
(1.71), yields:

𝑏𝑏𝑏𝑏𝑏𝑏 0.597 • 2500 bar


𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃 𝐷𝐷𝐷𝐷𝐷𝐷𝐷𝐷 � �= = 0.33
𝑚𝑚 100 • 3.052 • 14.53 m

Running a 30 m hose from the pump to the cutting nozzle would thus generate losses of
approximately 10 bar. It is interesting to note that in an earlier experimental study by Tursi
et. al. (1975) the measured pressure loss was slightly higher at 0.45 bar/m. The additional
losses are likely to have been incurred in the fittings along the line and/or by a change in
the line condition along its length.

Line Losses Often Exceed 50% of Energy Input


In practice, system line losses can be significant. Summers (1995) cites a Navy
study of a water jet system for cleaning ship boilers which, in operation, lost 71% of its
energy due to line losses. By optimizing system components, system losses were reduced
from 71% to 33%; still a significant loss by any measure but nevertheless a vast
improvement in system efficiency and performance.

102
Losses Also Occur When the Water Jet Exits the Nozzle
The work of a water jet is not completed when the jet leaves the nozzle. It is
completed only after the jet has impacted the target material. The deterioration of the water
jet while traveling across the stand-off distance represents the second source of losses.

Jet Energy Rapidly Dissipates With Stand-Off Distance


As noted in above sections, water jets are rapidly eroded both in size and pressure
as they travel away from the nozzle orifice. The erosion of power is significantly increased
when the jet in submerged and becomes a major design consideration at depths below 150
m, or its pressure equivalent.

Lowest Pressure and Largest Diameter Nozzle Offer Best Performance


For this reason, it is important to size the cutting jet to operate at the lowest pressure,
and the largest nozzle diameter consistent with the needs of the target ore mining operation.
Studies of the decay in water jet pressure away from the nozzle indicate that the pressure
loss is exponential with distance. (See Yanaida Approximation Equation (1.13) and Figure
18).

Impact Pressure as Function of Distance, Nozzle Size & Pressure


Summers (1995) reports a statistical regression on the pressure values measured, at
fixed distances between 15 and 75 cm, and under pressures of 280-700 bar with three
nozzle diameters (1.0, 1.25 and 1.5 mm diameters) to be of the order:

𝐼𝐼𝐼𝐼𝐼𝐼𝐼𝐼𝐼𝐼𝐼𝐼 𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃(𝑏𝑏𝑏𝑏𝑏𝑏) = 𝑐𝑐6 • 𝑑𝑑 2.93 • P 0.438 • 𝐷𝐷−0.939 (1.72)

Where:

𝑐𝑐6 Empirically determined constant


P Pressure of Water Jet (Bar)
𝑑𝑑 Nozzle diameter (in millimeters)
𝐷𝐷 Stand-Off Distance (in centimeters)

An alternate approximation of impact pressure for this range of system parameters is


presented by Summers (1995) as:

𝐼𝐼𝐼𝐼𝐼𝐼𝐼𝐼𝐼𝐼𝐼𝐼 𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃(𝑏𝑏𝑏𝑏𝑏𝑏) = 389 • 𝑒𝑒 −0.0165•𝐷𝐷 (1.73)

Where:

𝐷𝐷 Stand-Off Distance (in centimeters)

103
Jet Structure Can Be Enhanced By Sheathing
As the jet leaves the nozzle the pressure profile across the jet is akin to a step
function. There is a very rapid rise to full stagnation pressure. As the jet moves away the
profile changes. The outer sheath of the water is slowed by the surrounding air and this in
turn slows the next layer. Within a short distance the pressure profile is almost bell shaped
(See Figure 17). With further increases in distance the jet begins to break up into segments,
drops and fine mists. (See Figure 50 below). As distances increase further from the nozzle
the structure and corresponding cutting energy of the jet disperses rapidly. Jet structure
and effective cutting distances can be improved by sheathing the jet with a surrounding
spray, either of slower moving water or air.

Figure 50 – Water Jet Structure Deterioration With Increasing Distance (Summers 1995)

Cutting Effectiveness Limited to 200 Nozzle Diameters


As set out in Section 3.6, nozzle design is crucial to minimize jet losses once the
water exits from the nozzle orifice. As a general rule-of-thumb, a water jet’s cutting
effectiveness is limited to a maximum of 200 nozzle diameters from the orifice.

Dissipation (Read Losses) of Energy in a Case Study – A Cautionary Tale


Summers (1995) presents an interesting case study of the distribution of energy in
a prototypical water jetting system. It is equally applicable to an HBHM system and
highlights where (and how) energy losses arise within such a system. While it is for
illustration purposes only, the assumptions and parameters are typical of those found in
off-the-shelf equipment. It highlights the need for careful selection and integration of
system components to maximize water jet energy delivered to the target surface.

System Assumptions (This Section is Derived from Summers’ (1995) Example):


• 60 kW motor to power a high pressure pump
• Flow rate of 44 lpm

104
• Jet Pressure of 700 bar
• Flow passes through 10 m of 4.76 mm internal tubing into three nozzles
• Each nozzle orifice is 1.4 mm in diameter
• Stand-Off distance from nozzle to target surface is 30 cm
• Slot size being cut is 2.5 cm by 5 cm wide
• Nozzle traverse velocity of 60 m/min

Specific energy values can be calculated based on the amount of energy available at each
stage of the delivery line.

Stage 1 – Energy Input to the Motor


The volume of the material removed from the working surface each second is equal
to the depth of the cut multiplied by the width of the cut and the distance which the nozzles
have moved. Unless otherwise stated, all dimensions are assumed to be in millimeters
(mm). For each second, this can be calculated as follows:

Volume (v) = Depth of Cut (h) X Width of Cut (w) X Distance Moved (m) (1.74)

Where:
v volume of rock excavated in cu. mm
h depth of cut in mm
w width of cut in mm
m distance cut per sec in mm

Substituting the actual figures in this example gives:

Volume = 25 X 50 X 1000 = 1,250,000 cu. mm = 1,250 cu. cm (cc)

The motor power of 60 kW is equivalent to 60,000 joules, so that the overall


specific energy (i.e. the energy in joules required to remove 1 cc of rock) can be calculated
by simple division. The units most often reported are either Megajoules/cubic m or
joules/cc. These are numerically equivalent though we will use joules/cc unless otherwise
stated. From Equation (1.5), we know that:

Specific Energy = Energy/Volume Removed (1.5)

Or, by substituting the figures in this example,

Specific Energy = 60,000/1250 = 48 joules/cc

105
This is a critical metric as it considers the gross quantity of energy input into the
front-end of the system for which the user must pay versus the material recovered from the
back-end of the system for which the user is paid. The front-end energy input of 60,000
joules is also the gross starting figure from which downstream energy losses will be
deducted.

Stage 2 – Input Power to the Pump (90% of System Energy Input)

If the electric motor is assumed to be 90% efficient, the pump will receive only 54
kW (i.e. 90% X 60 kW) of power from the motor, or 54,000 joules. As the volume of rock
removed from the surface remains a constant, this lowers the Specific Energy calculation
as follows:

Specific Energy = 54,000/1250 = 43.3 joules/cc

The revised Specific Energy of 43.2 joules/cc is 90% of the theoretical maximum
Specific Energy calculation of 48 joules/cc and reflects that fact that only 90% of the input
energy of 60 kW is available to the pump or

Available Energy = Energy Available/Energy Input (1.75)

= 54,000 joules/60,000 joules = 90%

Stage 3 - Power Output from the Pump to the Water (86% of System Energy Input)

If the water delivered by the pump is flowing at a volume (Q) of 44 lpm and at a
pressure (P) of 700 bar at the pump exit, then the contained power can be calculated from
Equation (1.1)

E = ⦗(1.666•P) • Q⦘/1000 (1.1)

Where:

E Hydraulic Horsepower of Water Jet (kW)


P Pressure of Water Jet (Bar)
Q Flow Rate of Water Jet (liters/min)

Substituting the above figures, gives the following water power calculation:

E = ⦗(1.666•700) • 44⦘/1000 = 51.31 kW

106
Using the water power figure of 51.31 kW, yields the following revised Specific Energy
and Available Energy estimates:

Specific Energy = Energy/Volume Removed (1.5)

Specific Energy = 51,310/1250 = 41 joules/cc

Available Energy = Energy Available/Energy Input (1.75)

= 51,310 joules/60,000 joules = 86%

Stage 4 - Power Output to the Water Jet Nozzle Orifice (50% of System Energy Input)
Based on the above, the water has a power of 51,310 joules as it exits the pump.
We must now calculate the pressure losses in the line between the pump and the nozzle
orifice. In this example, it is assumed that the pressure loss in driving 44 lpm through 10
m of pipe with an internal diameter of 4.76 mm is roughly 26 bar for each m of tubing as
calculated from Equation (1.71) below:

𝑏𝑏𝑏𝑏𝑏𝑏
𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃 𝐷𝐷𝐷𝐷𝐷𝐷𝐷𝐷 ( 𝑚𝑚 ) = (0.597 • 𝑄𝑄 2 )/( 100 • 𝐷𝐷5 • 𝑅𝑅𝑅𝑅𝑅𝑅 0.25 ) (1.71)

Where:

Q Flow Rate of Water Jet (liters/min)


𝐷𝐷 Pipe diameter (cm)
𝑅𝑅𝑅𝑅𝑅𝑅 Reynolds number given by 1116.5•(Q/D)

From above, a flow rate Q of 44 lpm and pipe diameter D of 0.476 cm yields a Reynolds
Number for such a flow of (1116.5 X 44/0.476) or 103,205. Employing Equation (1.71),
yields:

𝑏𝑏𝑏𝑏𝑏𝑏 0.597 • 1936 bar


𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃 𝐷𝐷𝐷𝐷𝐷𝐷𝐷𝐷 � �= = 26
𝑚𝑚 100 • 0.0244436 • 17.924 m

In addition to the line losses of 26 bar per meter there is an assumed further loss of 35 bar
in turning the fluid flow 90 degrees from the borehole conduit to the nozzle. Thus the total
power loss in the water jet from leaving the pump to the point where it is discharged at the
nozzle can be calculated as follows:

107
Jet Pressure Line Loss = 10 m X 26 bar/m + 35 bar = 295 bar

Jet Pressure @ Nozzle Inlet = Input Pressure – Pressure Loss, or

Jet Pressure @ Nozzle Inlet = 700 bar – 295 bar = 405 bar

This yields the follow Power, Specific Energy and Available Energy calculations:

For Water Power from Equation (1.1):

E = ⦗(1.666•P) • Q⦘/1000 (1.1)

Where:

E Hydraulic Horsepower of Water Jet (kW)


P Pressure of Water Jet (Bar)
Q Flow Rate of Water Jet (liters/min)

We calculate the following:

E = ⦗(1.666•405) • 44⦘/1000 = 29.7 kW

Specific Energy = Energy/Volume Removed (1.5)

Specific Energy = 29,700/1250 = 24 joules/cc

Available Energy = Energy Available/Energy Input (1.75)

= 29,700 joules/60,000 joules = 50%

As shown above, about 50% of the system input energy is lost in getting the water from
the pump, through the line and out the nozzle.

Stage 5 – Power Delivered to the Target Surface (14% of System Energy Input)
If the fluid flowing to the target is separated into three nozzles, then at 44 lpm this
will require that each nozzle diameter be approximately 1.4 mm. (See Section 3.6,
Equations 1.3 and 1.31). The foregoing assumes a Coefficient of Discharge of 1.0 and the
jet pressure at the nozzle is 405 bar.

108
At a standoff distance of 30 cm, the Impact Pressure can be approximated by Equation
(1.73) below:

𝐼𝐼𝐼𝐼𝐼𝐼𝐼𝐼𝐼𝐼𝐼𝐼 𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃(𝑏𝑏𝑏𝑏𝑏𝑏) = 389 • 𝑒𝑒 −0.0165•𝐷𝐷 (1.73)

Where:

𝐷𝐷 Stand-Off Distance (in centimeters)

Substituting a stand-off distance of 30 cm gives an impact pressure at the surface of the


target of 237 bar.

For Water Power from Equation (1.1):

E = ⦗(1.666•P) • Q⦘/1000 (1.1)

Where:

E Hydraulic Horsepower of Water Jet (kW)


P Pressure of Water Jet (Bar)
Q Flow Rate of Water Jet (liters/min)

We calculate the following:

E = ⦗(1.666•237) • 44⦘/1000 = 17.4 kW

This water power estimate of 17.4 kW, however, is not a constant over the jet but as
illustrated in Figure 17 is a bell shaped curve. By approximating the curve to a triangular
function a simplified estimate of the amount of energy remaining in the jet can be made
and is estimated by Summers (1995) to be 8,100 joules. Recalculating the values for energy
efficiency result in the following:

Specific Energy = Energy/Volume Removed (1.5)

Specific Energy = 8,100/1250 = 6.5 joules/cc

Available Energy = Energy Available/Energy Input (1.75)

= 8,100 joules/60,000 joules = 14%

109
Targeted Rock Face Receives Only 14% of Energy Input
This is a rather shocking result which shows 86% of the system input energy being
dissipated (or lost) prior to the water jet impacting the target surface. Or stated differently,
only 14% of the input energy is available to excavate the rock face. While only 6.5 joules
of energy is required to remove one cc of rock, the system inefficiencies require 48 joules
of energy input to the HBHM system. The assumptions used in this case study by Summers
(1995) are typical of off-the-shelf equipment employed in HBHM and water jetting
applications. The case study underscores the importance of competent and thoughtful
system design and component selection to minimize energy losses and maximize the
energy available to impact the target surface.

Energy Losses Concentrated in Lines (36%) and Water Jet Post Nozzle (36%)
As highlighted in Figures 51 and 52 below, the largest system losses occur in the
form of: (i) line losses from the pump to the nozzle (36%); and (ii) jet energy dissipation
from the nozzle orifice to the target rock surface (a further 36%). These two components
of any HBHM system design should receive the greatest attention and scrutiny from the
technical team responsible for design and implementation of the HBHM system.

Stage 1 Stage 2 Stage 3 Stage 4 Stage 5


60K Joules 54K Joules 51K Joules 30K Joules 8
Joules
100% Energy 90% Energy 86% Energy 50% Energy 14%
Energy

Figure 51 – Energy Losses at Each Stage of HBHM System


(Summers 1995)

110
Energy Retained at Each System Stage (K Joules)
60
System Energy Retention (K Joules)
50

40

30

20

10

0
1 2 3 4 5
System Stage

Figure 52 – Energy Retention (K Joules) at Each Stage of HBHM System

3.15 Mine Unit Stability

Minimizing Subsidence & Cave-Ins


Stability of the overburden and roof rock in caverns is a major consideration in any
HBHM operation. The unintended collapse of a cavity could result in loss of mineral
reserves, surface subsidence, equipment loss, or worst, injury to personnel. Cavity stability
is a product of a complex set of multi-dimensional variables including in-situ stresses, rock
properties, cavity geometry, time and the rate and manner of excavation. Careful testing
of rock competence above the targeted zones of mining needs to be carried out prior to the
design of a mining plan. When mining is performed submerged, rather than at atmosphere,
charge water can be supplied to the annulus on the backside of the drill pipe as required to
maintain hydrostatic head on the well and cavity. A hydrostatic head of 150 meters in a
flooded cavern provides over 190 tons per square meter of average pressure on the rock
roof for support (Merlin 2012). Additional support can be introduced through injection of
water into the cavern on the annulus side of the drill pipe. In the case of atmospheric
mining, backfill may be added into caverns to ensure stability.

Mining Sequencing Can Help


The sequence of mining in certain instances may also be varied to minimize block
shifting or cave-ins. The standard ‘Bottom-Up’ mining sequence is where the cavern is
created and mined from the bottom most section upward. Because of varying rock
competencies a more ‘Top-Down’ approach may be justified whereby top sections are first
111
mined prior to casing and running deeper. Apart from cavity stability considerations,
varying the mining sequence also introduces the possibility of ‘high-grading’ thereby
improving financial returns.

Kimberlite Pipes Have Inherent Stability


In the case of kimberlite pipes, the ‘champagne flute’ shape of the deposit provides
some intrinsic stability. As the pipes have a small taper to them, this creates a gravity plug
and the overlying layers pressure-up the rock density to limit movement of the ceiling.
This feature of kimberlites further reduces the risk of collapse.

Critical Factors for Cavity Stability


There is relatively little published research on cavity stability. This, in large part,
reflects the complexity of modeling such situations given the multitude of dynamic and
interacting factors (e.g. in-situ stresses, rock properties, cavity geometry, etc.). Recent
work by Akbarzadeh (2014) for thin coal seams however yields some useful general
observations:

• Increasing internal pressure and cavity length will significantly increase the induced
maximum principal stress around the cavity.
• Depth of cover plays an important role, but not as significant as internal pressure and
cavity size.
• Applying negative pressure will cause the stress concentration to move from the cavity
roof to the side-wall and corner of the cavity.
• Increasing the horizontal stress will decrease the induced stress at the cavity roof corner.
• Reaching final optimum cavity size by using horizontal slides extraction is more stable
than using vertical slices.
• In cavities with small dimensions, the impact of internal pressures is less than those with
larger geometries.

Whether such observations hold for cavity stability in kimberlite deposits is yet to be
determined.

3.16 Infrastructure and Equipment

The Drilling Rig


An integrated HBHM system includes a drilling rig to drill and case (if required)
through the overburden section into the ore body. The rig can drill with conventional
annular circulation or airlift reverse circulation. The drill pipe is then tripped out after
drilling or bulk sampling and then the mining pipe and the jet miner is tripped in. The rig
must be modified to handle the tooling on the miner including specific safety requirements
and protections to handle the multiple types of fluids including high and low clean water,
112
annular charge water, return slurry water and both high and low pressure independent air
lines. Once the mining has transitioned from the pilot phase to full scale commercial
operations, the rigs can be separated into the roles of overburden drilling, mining and
backfilling (if required).

Rig Selection – Pipe Diameter and Weight Capacity Most Important Specs.
The priority considerations for rig selection are: depth rating, pipe diameter
handling capacity and rated weight capacity. Secondary considerations are productivity,
capacity to load large diameter equipment and suitability to handle the mining equipment.
The mining rig must be equipped to handle a multiport swivel with the necessary piping to
feed both the high pressure/high volume flow rate water and low and high pressure air
streams to the jet miner for the air shroud and the hydraulic lift system. Fortunately there
are a number of manufacturers who produce rigs with the foregoing capabilities.

Fluid Pressure & Volume Control – Twin Pumps Preferred


Pumping units provide fluids to the HBHM system for mining, for lifting the ore
and to separate the solids. Fluids need to be at adequate pressure and volume to cut into
the rock face and disaggregate it in a non-pulsating and non-surging constant flow. Once
cutting and lifting capacities have been defined from initial lab testing, they will need to
be adjusted during the pilot and mining phases. Due to varying rock consistencies, each
pass of the miner requires a slightly different parameter of pressure and volume to match
the distance cut. Pumps that deliver constant load over a range of pressures and volumes
settings are required. Generally, twin units are preferred for both high redundancy and
peak operating volumes.

Airlift
Utilization of airlift technology allows the pressure pumping capacity to be focused
on mining and to be delivered at the jet face. The recharge water supplied for lifting the
ore is pumped down the annulus of the well bore through the surface casing and outside
the drill pipe. Depending on the consistency of production and the density of the slurry
being lifted, a lower pressure high volume charge pump can be added at surface. This
pump can feed the ‘backside’ of the drill pipe through a rotating head seal at the top of the
casing. The additional pressure of the water pumped down the back side of the drill pipe
supplements the flow and allows the jet to cut in a cleaner environment by moving the cut
material out and reducing the density of the slurry above the intake at the bit. It is for this
specific reason that the HBHM system must have the ability to vary the distance between
the intake and the jet.

113
Air Compressors for Airlift & Shrouding
A large quantity of air is required for the hydraulic mining process in a submerged
environment. Suitable air compressors are available as off-the-shelf items for integration
into any HBHM system. The air is used to feed the air shrouding system for the hydraulic
jet down hole. This shrouding, as discussed earlier in the report, increases the cutting
distance of the jet by reducing the density of the water that the jet is traveling through. This
effectively increases net hydraulic horsepower farther out in the borehole, thus increasing
cavity diameter by up to 20%. Air is also needed for the hydraulic airlift system that
recovers ore to the surface. This hydraulic airlift system reduces the density of the fluid in
the mining pipe causing a vacuum to be created downhole, thereby lifting the slurry from
the rock face to the surface in a consistent manner.

Surface Separation and Screening


Separation and screening is required at surface as fluids are settled, separated and
re-circulated. This can be generally accomplished using conventionally sourced oilfield
‘linear motion shaker’ units. Hydro-cyclones are used when necessary to separate fines
from the post shaker slurry that has passed through the screens.
A typical equipment configuration for an HBHM operation is illustrated in Figure
53 below.

Figure 53 – Typical Equipment Configuration for HBHM Diamond Operation


(Merlin 2012)

114
3.17 Feasibility Study & Pilot Testing Requirements

Feasibility Study Requirements


Prior to the commencement of any HBHM operation, a full feasibility study (see
typical Table-of-Contents Below) is required to determine equipment requirements, mine
plan, production rates, processing requirements, economics, etc. The first step is ore
delineation to assess concentrations of the targeted metal (grade) and determine which
sections of the deposit are amenable to HBHM mining. In the case of kimberlites, there is
considerable variability of grade and specific gravity within the ore body. Past industry
experience has shown that the speed of volcanic flows affect the geotechnical properties of
kimberlite pipes not only with respect to height and depth of a section also in the cross
section of the kimberlite pipe itself. Care must be taken therefore in the exploration drilling
phase to adequately delineate both grades and densities by section.

HBHM FEASIBILITY STUDY – KEY COMPONENTS

1.0 Introduction
2.0 Tenements & Permitting
3.0 Geology & Mineral Resources
4.0 Geotechnical Analysis
5.0 Mining
6.0 Metallurgy
7.0 Process Plant Design
8.0 Infrastructure
9.0 Environmental
10.0 Project Economics
11.0 Recommendations for Pilot Test

Grade and Cutting Pressure are the Key Parameters


Once the kimberlite pipe has been extensively drilled and sampled to determine
diamond grades by section, it is then necessary to assess the threshold and cutting pressure
of each section of the ore body. The threshold pressure is the pressure required to push the
water behind the rock grains and to disaggregate the rock. The cutting pressure is that
pressure required to disaggregate the ore at a fast enough rate, and a low enough
consumption of overall energy and cost, to render the operation economic. As a practical
rule, the cutting pressure is typically 3-4 times the threshold pressure. Based on a balanced
system configuration, and currently available equipment in the commercial market, the
limitation of cutting pressure is 5,000 psig. Threshold and cutting pressures are estimated
by laboratory testing of core samples, typically contracted through a university with such

115
specialized facilities (e.g. Colorado School of Mines or Missouri University of Science &
Technology).

Pilot Test – A Necessary Pre-Production Step


Following completion of a positive feasibility study, a pilot test will be required to
better define the equipment and procedures to successfully mine the targeted ore body, to
efficiently deliver ore to the surface, operating costs, production rates, etc. The pilot test
will assist in determining optimal jet pressures, flow rates and velocities, lifting recovery
densities and velocities and whether HBHM can mine the targeted ore sections at an
economic rate. Only upon completion of a successful HBHM pilot test can the operator
consider commencement of commercial scale operations. Key information to be
determined from the field pilot test includes the following:

• Threshold and cutting pressures


• The effective jet pressures and jet engineering at given diameters
• The optimum flow rate of fluids and air at the jet
• The jet traverse velocity
• The vertical spacing of cuts to minimize roof failures
• The lifting, recovery density and velocity
• The extent of diamond damages incurred

Initial Parameters to Be Estimated By Laboratory Testing


As noted above, the initial threshold and cutting pressures will be determined by
laboratory water jet testing of core sections. Similarly, the initial estimates of jet traverse
velocity, vertical spacing of the cuts and flow rates will be determined from laboratory
surface testing of larger blocks of ore. The foregoing estimates will then provide sufficient
input data to render initial estimates of slurry densities and lifting speeds.

116
Chapter 4
Borehole Mining for Diamondiferous Deposits

4.1 HBHM Applied to Diamondiferous Deposits

Historic HBHM Testing – Never Applied to a Kimberlite Pipe


As noted in Section 3.1, HBHM technology has been pilot tested the past 40 years
on deposits of bauxite, coal, gold placers, iron ore, manganese, oil sands, phosphates,
titanium/zircon mineral sands and uranium with mixed results. While the technology was
used successfully in the mid-1990s for bulk diamond sampling on a variety of Canadian
kimberlite deposits it has never been commercially applied to the mining of
diamondiferous kimberlites. This is surprising as such deposits would seem to be natural
targets given their vertical geometry, high value, typically small size and heterogeneous
ore types which often include sections of soft ore that are easily fragmented.

Figure 54 – Idealized Kimberlite Pipe Showing Economic Limits of Open Pit and Extended Reach of Jet
Mining
(Source: www.gia.edu)

Conceptually, HBHM is Well Suited to a Kimberlite’s Geometry, Size & Grade


Figure 54 above illustrates an idealized kimberlite pipe with the economic limits of
its extractable open pit ore highlighted in green as well as the deeper ore zones accessible
via HBHM mining methods highlighted in blue.
In Figure 55 below a single borehole is drilled into the deeper zones beneath the
already mined out open pit of this idealized kimberlite pipe. The vertical borehole
orientation is an excellent match to the narrow, vertical geometry of the typical kimberlite.

117
Figure 55 – Idealized HBHM Hole In Kimberlite Pipe
(Merlin 2012)

HBHM’s Ability to Selectively Mine is a Key Advantage


Figure 56 below continues this sequence of illustrations by highlighting selected
sections (highlighted in red) of individual boreholes drilled into the deeper zones targeted
for HBHM mining and extraction. The highlighted sections in the deeper zones are deemed
amenable to economic extraction via HBHM methods based on minimum diamond grade
and cutting threshold pressure parameters developed during the feasibility study and pilot
testing phases. Only these highlighted sections have a sufficiently high concentration of
diamonds and low cutting pressure thresholds to be amenable to economic extraction via
HBHM methods. All other sections will be left intact as their diamond concentration is
too low and/or their kimberlite ore is too hard to cut via a water jet.

118
Figure 56 – Idealized HBHM Holes in Kimberlite Pipe with Targeted Sections Highlighted in Red
(Merlin 2012)

HBHM Has Potential to Unlock Value from Stranded Pipes


The above series of illustrations (i.e. Figures 54-56) demonstrate conceptually the
merits of HBHM mining of kimberlite pipes. HBHM is the only mining methodology that
will allow highly selective mining of deeper, higher grade and weathered sections of a
kimberlite pipe in an economic manner. It has the potential to unlock economic value
from what to date have been uneconomic, stranded pipes, particularly in Africa.

Merlin Diamonds – HBHM Pilot Test


The first (and only) publicized effort to test HBHM methods on a kimberlite pipe
is by Australian-based Merlin Diamonds Ltd. In 2012, Merlin engaged Kinley Exploration
LLC and their Australian affiliate (Jet Mining Ltd.) to undertake a feasibility study and
pilot test of HBHM on their Merlin cluster of kimberlites. Merlin is the owner of the Merlin
diamond mine located in the Northern Territory of Australia. It comprises 14 kimberlite
pipes, grouped in four clusters. Nine of these pipes were subject to open-pit mining by Rio
Tinto over a five year period commencing in 1998. The operations ceased in 2003 when
all the ore accessible by conventional surface mining was exhausted. Kinley prepared a

119
feasibility study in late 2012 confirming the economic viability of HBHM technology to
selectively mine deeper zones of interest. Trial mining commenced in August 2013. Early
results were promising with an announcement by Merlin on 17 October 2013 that read, in
part, as follows:

“Merlin Diamonds Ltd would like to announce an update to its trial borehole
mining operations at the Merlin Diamond Mine in the Northern Territory. The borehole
mining rig has been operating for over one month and has achieved success in a number
of key areas. The hydraulic jetting tool that cuts the kimberlite material at depth has been
proven to effectively cut the weathered kimberlite and is able to produce diamond bearing
ore suitable for lifting via the mining rods. The hydraulic lifting system has been proven to
lift material to the surface of the pit and is able deliver ore to the shaker screen located on
the ground surface adjacent to the pit. The processing plant has also achieved nameplate
capacity of 75t per hour.

Production rates experienced from the borehole mining technique require


optimising to guarantee maximum recovery and profitability through mining, and full
utilisation of the processing plant capabilities. The Company is now at a stage to complete
engineering work that will increase this rate of production. Production rates through the
borehole mining technique have been attained at other diamond deposits and Merlin’s
engineering and feasibility team will spend the coming months to complete the necessary
work. Given the above the Company has decided to enter into a production hiatus until this
engineering work is complete to ensure maximum shareholder return. “

Merlin’s Mining Plan


Merlin plans to begin mining in the initial passes without the air shroud as the
kimberlite face will be very close to the jet nozzle. The tool will be rotated and the jet will
be used to put the weathered kimberlite into slurry to be carried to the surface separation
equipment. As dictated by mining conditions, the air shroud will be engaged to ensure
maximum cutting distances are achieved. As the cavity is mined, the tool will be lowered
to the next level by rotating ahead in the predrilled hole with the bit on the bottom of the
hole, repeating the hydraulic borehole mining process at the next level. When the targeted
zone has been mined to the bottom of the weathered kimberlite layer, the HBHM
equipment will be removed and mobilized to the next location. Merlin’s planned mining
pattern for one of their kimberlite pipes, along with targeted sections, is illustrated in
Figures 57 and 58 below.

120
Figure 57 –HBHM Mining Pattern
(Merlin 2012)

Figure 58 –Cross Section of Kimberlite Pipe With HBHM Sections Highlighted


(Merlin 2012)

121
Merlin’s Initial Cutting Parameters

Merlin’s lab testing of its kimberlite ore suggests that ore sections with a specific
gravity of more than 2.3 g/cc will be too hard to cut by water jet and therefore not
amendable to HBHM.
Initial parameters for cutting of the weathered kimberlite sections have been
determined by Merlin to be 760 to 1,325 lpm (200-350 gpm) at a pump pressure of 138 to
345 bar (2,000-5,000 psi). This combination is deemed to deliver enough water flow and
pressure to cause a significant impact at the face of the weathered kimberlite to cause it to
turn to slurry with solids small enough to return up the airlift tract and be recovered through
the hydraulic borehole mining equipment to surface.
Merlin will employ a one cm nozzle and is estimating an initial traverse velocity of
10 inches/sec. at the start of mining. This will deliver a tight water jet with high impact
pressure. Once the cavity is opened up, Merlin anticipates that the traverse velocity will
be increased to the up to 60 inches per second with vertical spacing of 4 to 10 inches
between cuts. The vertical spacing will correspond directly to the distance between the
nozzle and the cut.

Results Unknown
No further updates have been provided as of the date of this report. Final results
of the trial mining exercise will be of great use in further assessing the viability of HBHM
technology in general and for its potential applicability to the selective mining of
diamondiferous deposits.

4.2 Geological & Mining Parameters for Success

HBHM Offers Great Promise for Kimberlite Pipes


For the many reasons cited earlier in this thesis, we believe that the use of HBHM
technology on kimberlite pipes warrants further industry investigation. The incentives for
success are further magnified by the projected shortages of diamonds and lack of recent
exploration success in discovering new economic kimberlite deposits.

Favorable Mining & Geologic Characteristics


Mining and geologic characteristics of kimberlite deposits which lend themselves
favorably to HBHM mining methods may be summarized as follows:

Favorable Kimberlite Geological &


Mining Characteristics For HBHM Application

Thickness of Targeted Zone: >10 meters

122
Width of Targeted Zone: Full pipe width
Min. Tonnage: > 5,000,000 tonnes of HBHM ore
Grade (Carats Per Hundred Tonnes): >30 carat per hundred tonnes (‘CPHT’)
Avg. Value Per Carat: > US$300/Carat
Specific Gravity: < 2.00
Depth: 50-250 Meters
Water Source: Local Water Source Highly Desirable

As noted above, a number of geological and mining characteristics must be present,


a priori, to render a kimberlite pipe a favorable candidate for HBHM methods. First, the
minimum thickness of a targeted section of the pipe should be not less than 10 meters and
ideally should span the full cross section of the pipe. The zone should not be less than 50
meters and not than 250 meters from the drilling surface. The ore in the targeted section
should be relatively soft with a SG of not more than 2.00. The grade of the targeted zone
should be at least 30 carats per hundred tonnes (‘CPHT’, the standard industry measure of
grade) and the average value per carat should greater than US$300. Finally, the aggregate
tonnage across all targeted zones should be at least 5,000,000 tonnes and there should be a
readily available supply of water. With all the foregoing factors present, the likelihood of
success of HBHM will be greatly enhanced.

4.3 Targeted Diamondiferous Deposits

Africa Offers the Most Opportunities


Diamondiferous kimberlites are found throughout the world though Africa, Russia,
Canada and Australia possess a disproportionate number of economic deposits. Africa has
by far the largest number of documented diamondiferous kimberlite occurrences. See
Figure 59 below.

123
Figure 59 – African Primary and Secondary Diamondiferous Deposits
(Source: www.diamondstamps.eu)

Many Targets Identified


Africa has 54% of known diamond resources placing it significantly ahead of
Russia, the second largest source of diamond resources with 14%. From public records,
we have compiled a list prospective HBHM test pipes held by financially distressed private
and small publicly listed companies. Information is hard to source as there is no published
directory of kimberlite pipes. A summary table of prospective worldwide targets identified
in our search is set out below.

PROJECT
COUNTRY OWNER NAME STAGE RESOURCE
Cue & Soopy
Australia Sunrise Resources Bore Adv. Expl. N.A.
Australia Merlin Diamonds Merlin Cluster Trial Mining 30 MT @ 24 CPHT
Australia Kimberley Diamonds Ellendale Care & Maint. 97 MT @ 4 CPHT
Botswana Firestone Diamonds BK11 Pipe Care & Maint. 12MT @ 9 CPHT
Botswana Kimberley Diamonds Lerala Prod. 20MT @ 27 CPHT
Brazil Minerado Bravo Cavado Canastra 1 Adv. Expl. 10MT @ 18 CPHT
Canada Shear Diamonds Jericho Pipe Care & Maint. 3.4MT @ 90 CPHT
Canada Metalex U2 Pipe Adv. Expl. 41MT @ 30 CPHT
Canada Dianor Resources Inc. Leadbetter Pipe Adv. Expl. 566MT @ 28 CPHT
Guinea Stellar Diamonds Droujba Bulk Sampling 4.5MT @ 88 CPHT
Guinea Stellar Diamonds Baoule Bulk Sampling 22 MT @ 30 CPHT
Lesotho Storm Mountain Diamonds Kao Mining 150 MT @ 7 CPHT
Lesotho Paragon Diamonds Lemphane Adv. Expl. 49 MT @ 2 CPHT
Lesotho Paragon Diamonds Motete Adv. Expl. 1.6 MT @ 65 CPHT
Sierra Leone Octahedra Holdings Pipe 3 Adv. Expl. 10 MT @ 30 CPHT

124
Sierra Leone Stellar Diamonds Tongo Adv. Expl. 0.6 MT @ 180 CPHT
South Africa Blue Rock Diamonds Kareevlei Cluster Adv. Expl. 8 MT @ 4.5 CPHT
South Africa Diamond Corp. Lace Development 35 MT @ 40 CPHT
Tanzania Stanton Ltd. Mahene Pipe Adv. Expl. 27 MT @ 8 CPHT
Tanzania Stanton Ltd. Itanana Pipe Adv. Expl. 10 MT @ 3 CPHT

* CPHT = Carats Per Hundred Tonnes

Table 5 – Diamond Exploration Targets

Targets Should Be Approached for Joint Venture


There are a number of potential kimberlite pipes that have favorable mining and
geological characteristics for the application of HBHM technology. They are highlighted
in Table 5 above and as previously noted the preponderance are located in Africa. They
range from projects at an advanced exploration stage to those on care-and-maintenance as
the prior operations were uneconomic using conventional mining methods. Nearly all of
the above noted projects are held by entities that are in a state of financial distress and
likely would be receptive to an approach by a party offering to fund an HBHM pilot study
in exchange for a negotiated equity interest in their kimberlite pipe. As a next step, we
would propose seeking financial backing for such a pilot test and once financing is secured
enter into negotiations with the above noted companies for a joint venture.

125
Chapter 5
Conceptual Economics

5.1 Conceptual Economic Model

No Industry Data So We Developed Our Own Economic Model


Given the lack of industry experience with HBHM operations, there is limited data
available for estimating capital and operating costs. As a substitute, we have developed a
financial model of a conceptual HBHM mining operation applied to an idealized kimberlite
pipe. While admittedly an imperfect approach, it represents a useful starting point for
assessing the potential economics of HBHM technology as applied to a kimberlite diamond
mining operation.

Ore Body Parameters


Our ‘idealized’ kimberlite pipe has the following characteristics.

Recoverable Tonnage: 5,500,000 tonnes


Grade: 30 CPHT
Specific Gravity: 2.00
Avg. Value Per Carat: US$350

Mining & Process Parameters

Our conceptual mining operation is based on a two rigs operating on a continuous


basis. This is largely modeled on the proposed Merlin diamond operation in Australia and
has the following parameters:

Process Plant Recovery Rate: 90%


Production Rate: 70 tonnes of ore per hour
Operating Hours: 24 Hours Per Day
Availability: 85%
Production Per Year: 521,220 tonnes per annum
Avg. Slurry Density: 22% solids, by mass
Slurry Recovery Rate: 1250 GPM

Operating Costs
This gives rise to the following operating cost estimates (again per Merlin’s data):

Operating Cost Per Tonne Ore: US$30/tonne

126
Capital Costs
Initial Capital costs for a two rig mining configuration are estimated to be on the
order of US$22 million as follows:

Jet Mining System (x 2): US$ 9,000,000


DMS Plant: US$ 4,000,000
Pumping Units (x2) : US$ 1,000,000
Instrumentation (x2): US$ 400,000
Shaker Tables: US$ 200,000
Piping, Valves: US$ 400,000
Misc. Equipment: US$ 1,000,000
Camp Infrastructure: US$ 2,500,000
Power: US$ 1,500,000
Mob. & Project Set-Up: US$ 2,000,000
Initial Capital Costs US$22,000,000

Sustaining capital is estimated to be 20% of the initial capital costs, or US$4.4


million per annum.

5.2 Anticipated Returns

Anticipated HBHM IRRs Very Attractive @ 79%


A financial model has been constructed in Microsoft Excel using the above noted
parameters. The operation has a mine life of 10.5 years and generates a highly attractive
IRR of 79% and NPV (@ 10% discount rate) of US$80 million. An excerpt of the financial
projections is as set out below. Sensitivity analysis shows that the grade of the deposit
could be reduced from 30 CPHT to as low as 18 CPHT while still yielding an acceptable
of IRR of 17%.

BOREHOLE MINING - CONCEPTUAL FINANCIAL MODEL

YEAR 0 YEAR 1 YEAR 2 YEAR 3


PRODUCTION (TONNES) 521220 521220 521220
GRADE (CPHT) 30 30 30
PROCESS RECOVERY RATE 90% 90% 90%
CARATS PRODUCED 140729 140729 140729
PRICE PER CARAT $350 $350 $350

REVENUES $49,255,290 $49,255,290 $49,255,290

OPERATING COST/TONNE $30 $30 $30

127
TOTAL OPERATING COSTS $15,636,600 $15,636,600 $15,636,600
NET PROFIT BEFORE TAX $33,618,690 $33,618,690 $33,618,690
INCOME TAX (@ 35%) $11,766,542 $11,766,542 $11,766,542
NET INCOME AFTER TAX $21,852,149 $21,852,149 $21,852,149

CAPITAL EXPENDITURES -$22,000,000 -$4,400,000 -$4,400,000 -$4,400,000

NET CASH FLOW -$22,000,000 $17,452,149 $17,452,149 $17,452,149

NPV (@10%): $79,929,808


IRR 79%

Table 6 –Conceptual Financial Model

5.3 Key Takeaways From The Financial Analysis

Key Takeaways - HBHM Is A Potential Game Changer


Key takeaways from the conceptual HBHM economic model are as follows:
• Capital costs for an HBHM operation are significantly lower than for a comparable
conventional diamond kimberlite operation. In the conceptual case, initial capital
costs were estimated at US$22 million for a two rig configuration which mines and
processes approximately 500,000 tonnes per annum of ore yielding production of
approximately 141,000 carats per annum.
• Attractive financial rates of return can be achieved for relatively modest scaled
operations (NPV-US$80 million; IRR-79%). Such small scale operations (i.e.
500,000 tonnes per annum of ore throughput) would be sub-economic by
conventional mining methods.
• Economic rates of returns are achievable for low grade deposits. Sensitivity
analysis suggests that an attractive IRR of 17% can be achieved with a grade as low
as 18 CPHT (vs. Base Case of 30 CPHT). This grade would likely be uneconomic
for even large-scale conventional operations.
• The ability of HBHM technology to render economic smaller sized and lower grade
kimberlite deposits represents a unique business opportunity to exploit such
‘stranded’ pipes and unlock their intrinsic value.

128
Chapter 6
Conclusions, Action Plan & Challenges

6.1 Conclusions

HBHM Offers Low Cost Alternative to Conventional Mining


Conceptually, HBHM offers a low cost alternative to conventional mining of
kimberlite pipes. It is ideally suited to exploit these typically modest, vertically oriented
ore bodies that are often not economic to exploit using conventional mining methods. The
technology promises significantly lower capital and operating costs than conventional
mining. Overburden does not have to be removed, underground infrastructure is not
required and massive fleets of earth moving equipment are not required. Production is
immediate, highly selective and in the form of a slurry which further reduces transport and
processing costs. Groundwater sources and flows are not disturbed. Environmental
impacts are greatly diminished.

HBHM is a Potential Game Changer


As demonstrated in the conceptual economic model, HBHM is a potential game
changer for the exploitation of kimberlite pipes which are not amenable to conventional
mining methods due to sub-economic size, low grade or remote location. Capital costs
are an order of magnitude lower than those required in conventional mining approaches.
A 500,000 tonne per annum operation requires only US$22 million of development capital
vs. US$150+ million for a comparably sized conventional operation (e.g. Firestone’s
Liqhobong – US$222 Million or Lucara’s AK6 Karowe – US$150+ Million). Financial
returns are highly attractive (IRR’s of 50-100%). The technology unlocks value in smaller
or lower grade diamondiferous pipes which are otherwise uneconomic using conventional
mining methods. Minimal surface disturbance and infrastructure requirements along with
the capacity to operate in a water saturated environment are added benefits of HBHM
technology.

6.2 Challenges

Airlift Component May Be ‘Achilles Heel’


The principal challenge is to demonstrate that HBHM technology can be advanced
beyond the prototype stage to full commercial scale whether on kimberlites or other ore
types. While various pilot tests on a spectrum of ore types undertaken the past forty years
have been positive, none have demonstrated that HBHM technology is sufficiently robust
to achieve and sustain commercial levels of production over long periods of time. The
weakness of the system is likely to be less in the mining of the rock face than in the
consistent and continuous lifting of the slurry from downhole to the surface. This may be

129
the ‘Achilles heel’ of the technology. Notwithstanding this potentially fatal flaw, the
economic and environmental benefits of the technology are sufficiently promising to merit
further industry attention and effort.

6.3 Action Plan

Targeted Companies Should Be Approached For Joint Venture


A number of kimberlites have been identified as potential HBHM candidates. They
are owned by financially distressed private and small market capitalization publicly traded
diamond explorers who would likely be receptive to an offer to undertake a pilot test in
exchange for a significant ownership interest in their kimberlite pipe. It is the intent of the
author to seek financial backing for the concept, and if successful, initiate discussions with
the owners of the targeted pipes. If a successful joint venture can be negotiated, a HBHM
feasibility study would be undertaken followed by a pilot test to validate the assumptions
of the feasibility study. Full scale commercial operations could commence if the pilot test
results are favorable.

6.4 Areas of Future Research

Field Data is Lacking


The U.S. Bureau of Mines funded a number of significant (and successful) field
trials of hydraulic bore hole mining during the 1970s and 1980s. The Clinton
Administration’s shuttering of the U.S. Bureau of Mines in 1996 brought this research
initiative to a halt. Since this time, there have been scattered field tests by private
companies but the efforts have been isolated and in most cases results have not be made
available to the public. Building on the early progress of borehole mining will require first
and foremost renewed industry commitment to field trials of the technology. In the absence
of such field trials, progress will be muted and borehole mining will remain a prototype
technology.

Enhanced Predictive Models of Rock Cutting Required


A major challenge in characterizing and modeling the failure mechanisms for rock
subjected to a water jet is the fact that rock is discontinuous, inhomogeneous, anisotropic
and inelastic. This is in sharp contrast to the ‘ideal’ material for analytical models which
is assumed to be continuous, homogeneous, isotropic and linearly elastic. The presence of
fluids adds to the complexity. As a consequence, analytical models cannot be relied upon
to provide the necessary predictive results for rock failure by water jet. Given the
intractable nature of the problem, there is increasing reliance on numerical models to
simulate and predict rock failure and cutting performance. While much progress has been
made of late, more efficient and robust predictive models of rock cutting by water jets are

130
needed. The current models either have unrealistic data input requirements or are overly
simplistic and consequently do not provide useful predictions. At present, the best
predictive water jetting model is still laboratory testing of representative rock specimens
coupled with blind hope that the lab results will approximate field performance. We refer
to this as the “spray-and-pray” approach. Key to advancing the current state of the art is a
better understanding of the conditions that govern fracture initiation, growth and the
pressurization of such fractures by the water jet fluid.

Abrasives Offer Great Potential but Need More Research


The introduction of abrasives into the water jet stream offers the potential to greatly
expand the range of ores that can be cut by HBHM, particularly for harder ore types like
granite, kimberlites, etc. However, there is little or no field experience with the use of
hydro-abrasive jets in HBHM applications. The research literature is equally vacant. This
represents a fruitful area of future theoretical and empirical research.

Optimizing Airlift Performance


Like rock failure, the theoretical study of airlift performance is complex and
dependent on numerous factors including, but not limited to, fluid-solid mix type, flow
regime, fluid density, pipe diameter, pipe length, air flow rate and pressure, fluid flow rate
and pressure, submergence ratio, boundary layer considerations, etc. Better predictive
models are needed to ensure that selected airlift design parameters will yield the desired
field performance. There has been considerable theoretical and empirical work undertaken
by past researchers but much more needs to be done to advance this critical component of
a borehole mining system, including evaluation of alternative pumping technologies (e.g.
down hole mechanical pumps).

Nozzle Design and Performance


Nozzle design is critical to the rock cutting performance of the water jet. There has
been virtually no research on HBHM nozzle design since 1976 when the USBM funded a
study of optimal nozzle design for mineral excavation (Lohn and Brent, 1976). It seems
probable that further improvements in nozzle design could be achieved by integrating
recent advances in computational fluid dynamics, materials science, control systems, etc.
In addition, since cutting power dissipates rapidly with increasing stand-off distance from
the nozzle, consideration should also be given to mechanical innovations such as a
‘telescopic nozzle’ which would physically extend outward as the cavern size expands to
minimize stand-off distance. All the foregoing, in our opinion, could yield significant
improvement in the cutting performance of HBHM systems.

131
Pulsating & Cavitation Water Jets Holds Great Promise
The fact that pulsation and cavitation water jets allow rock to be cut at a fluid
pressure significantly below the stagnation pressure of a continuous high pressure jet offers
great promise. This is one of the most promising areas of future research, particularly at
significantly higher flow rates than achieved to date and over a wider spectrum of minerals
(since the destructive pressures generated by pulsation and cavitation exceed those
necessary to disintegrate even the strongest of natural minerals).

132
REFERENCES

Abulnaga, B., (2002), Slurry Systems Handbook, McGraw-Hill Handbooks

Agus M, Bortolussi A., Ciccu R., Kim W., and Manca P.P., (1993), The Influence of Rock
Properties on Waterjet Performance, 7th International American Water Jet Conference,
Seattle WA, pp. 427-442

Akbarzadeh, Y., (2014), Protocols for Evaluating The Stability of Induced Cavities in
Underground Thin Seam Coal Deposits Created By In-Situ Borehole Mining, PhD Thesis,
Colorado School of Mines

Alehossein H. and Boland J.N., (2004), Strength, Damage and Fatigue of Rock, SIF2004,
Structural Integrity and Fracture

Arens V. Zh. And Khrulev A.S., (2003), Experience of Hydraulic Mining From Deep-
Seated Gold Placers, Vol. 39, No. 1, Journal of Mining Science

Bain & Company, Inc., (2011), Diamond Industry Report

Bain & Company, Inc., (2013), The Global Diamond Report

Benjamin T. B. and Ellis A.T. (1966), The Collapse of Cavitation Bubbles, Philosophical
Transactions of the Royal Society of London, Series A, Mathematical and Physical
Sciences, Vol. 260, No. 1110, pp. 221-240

Bieniawski Z.T., (1967), Mechanism of Brittle Fracture of Rock: Part I – Theory of the
Fracture Process, Vol. 4, Int. J. Rock Mechanics and Mining Sciences and Geomechanics,
Abstracts, pp. 395-406

Black Range Minerals, (13 February 2012), ASX Press Release, Positive Preliminary
Evaluation at the Hansen Uranium Deposit, 13 February 2012

Brook N. and Summers N.A., (1969), The Penetration of Rock By High Speed Water Jets,
Int. J. Rock Mech. Min. Sci., Vol. 6, pp. 249-258

Brunton J.H. and Rochester M.C., (1979), Erosion of Solid Surfaces by the Impact of Liquid
Drops, Erosion-Treatise on Materials Science and Technology, C.M. Preece ed., pp. 185-
248

133
Chahine G.L., Johnson, Jr. V.E. and Frederick G.S., (1982), The Feasibility of Passively
Interrupting Water-Jets for Rock Cutting Applications, Hydronautics, Inc. Technical
Report 8228-1, p. 75

Cheung J.B., Hurlburt, G.H., Scott, L.E. and Veenhuizen, S.D., (Sept. 1976), Application
of a Hydraulic Borehole Mining Apparatus in the Remote Extraction of Coal, Flow
Industries, Inc. Kent, Washington, Prepared for the United States Bureau of Mines

Crow, S.C, A Theory of Hydraulic Rock Cutting, (1973), International Journal of Rock
Mechanics and Mining Science, Vol. 10, pp. 567-584

Davis L.L., (March 1981), Bureau of Mines Research to Improve Underground


Metal/Nonmetal Mining Technology, Mining Engineering, pp. 305-312

Dimitrijevic B., Pinka J. and Mitrovic, V., (2004), Selection of Technological Parameters
in Borehole Mining Production by Technical Deep Drilling and Hydroexploitation, Vol.
9, Acta Montanistica Slovaca. Pp. 160-167

El-Saie A. A., (1977) Investigation of Rock Slotting by High Pressure Water Jet for use in
Tunneling, Doctoral Dissertation, Mining Engineering, University of Missouri-Rolla

Engin I.C., (Sept. 2012), A Correlation for Predicting the Abrasive Water Jet Cutting
Depth for Natural Stones, S. African Journal of Science, Vol. 108 (9/10)

Evans, A.G., 1979, Impact Damage Mechanics: Solid Projectiles, Erosion Treatise on
Materials Science and Technology, Vol. 16, C.M. Preece ed., Academic Press

Evers J., (October 1984), Update on Hydraulic Mining in the U.S., Mining Engineering,
pp. 1418-1421

Farmer I.W. and Attewell P.B., (July 1965), Rock Penetration by High Velocity Water Jets,
International Journal of Rock Mechanics and Mining Science, Vol. 2, pp. 135-153

Fowkes R. S. and Wallace J.J, (1968), Hydraulic Coal Mining Research, Assessment of
Parameters Affecting the Cutting Rate of Bituminous Coal, Report of Investigations 7090,
Bureau of Mines, Department of the Interior

134
Fujimoto, H., Murakami, S., Omura, A., and Takuda, H., (2004), Effect of Local Pipe Bends
on Pump Performance of a Small Airlift System in Transporting Solid Particles, Int. J. Heat
Fluid Flow, Vo. 25, No. 6, pp 996-1005

Fyall A.A. (1967), Second Rain Erosion Conference, Lake Constance, Germany, p. 428

Giot, M. (1982), Three-Phase Flow, Handbook of Multiphase Systems, G. Hetsroni, ed.,


paras. 7.2 and 7.2.2.8, New York, McGraw-Hill

Griffith A. A., (1920), The Phenomena of Rupture and Flow in Solids, Phils Trans., A221.
Pp. 163-198

Griffith A. A., (1924), Theory of Rupture, Proceedings of the First International Congress
for Applied Mechanics, pp. 53-64, J. Waltman Jr. Press

Harris H.D., Mellor M., (1974), Cutting Rock with Water Jets, International Journal of
Rock Mech. Min. Sci. and Geomech., Vol. 11, pp. 343-358

Hashish, M., 1984, A Modeling System of Metal Cutting with Abrasive Waterjets,
Transactions of the ASME, Vol. 106, pp. 88-100

Hashish, M., 1991, Abrasive Jets, Section 4 – Fluid Jet Technology-Fundamentals and
Applications, Waterjet Technology Association, St. Louis, MO

Hertz, H. (1882), Ueber die Beruhrung Fester Elastischer Korper, Journal fur die Reine
und Angewandte Mathematik, Jg. 92, pp. 156-171

Heymann F.J., (1970), Toward Quantitative Prediction of Liquid Impact Erosion, ASTM
STP 474, American Society for Testing and Materials, pp. 212-248

Hood M., Nordladn R and Thimons E., (1990), A Study of Rock Erosion Using High
Pressure Water Jets, International Journal Rock Mech. Min. Sci and Geomech., Vol. 27,
pp. 77-86

Hrabik J. and Godesky D, (1983), Economic Evaluation of Borehole and Conventional


Mining Systems in Phosphate Deposits, United States Bureau of Mines Information
Circular

Hrabik J.A, (1986), Economic and Environmental Comparison: Borehole Mining versus
Conventional Mining of Phosphate, January Mining Engineering, pp. 33-39

135
Hu, D., Tang, C.L., and Liao, Z.F., (2012), Modeling and Analysis of Airlift Pumping
System Based on the Momentum Theorem, J. Drain. Irrig. Mach. Eng., Vol. 30, No. 5, pp.
592-597

Hwang J.B.G. and Hammitt F.G., (1977), High-Speed Impact Between Curved Liquid
Surface and Rigid Flat Surface, J. Fluids Eng. 99(2), pp. 396-404

Jacobs, B.E.A., (1991), Design of Slurry Transport Systems, Elsevier Science Publishers

Jaeger, J.C., Cook N.G.W., and Zimmerman, R.W., (2007), Fundamentals of Rock
Mechanics, 4th Edition, Blackwell Publishing

Jeng F.S., Huang T.H. and Hilmersson S., (2004), New Developments of Watejet
Technology for Tunnel Excavation Purposes, Tunneling and Underground Space
Technology, Vol. 19, pp. 438-445

Kassab, S.Z., Kandil, H.A., Warda, H.A. and Ahmed, W.H., (2007), Experimental and
Analytical Investigations of Airlift Pumps Operating in Three-Phase Flow, Chem. Eng. J.,
Vol. 131, pp. 273-281

Kassab, S.Z., Kandil, H.A., Warda, H.A., Ahmed, W.H., (2008), Air-lift Pump
Characteristics Under Two-Phase Flow Conditions, Int’l J. Heat Fluid Flow

Kato, H., Miyazawa, T., Tiyama, S., and Iwasaki, T., (1975), A Study of and Airlift Pump
for Solid Particles, Bull. ISME, Vol. 18, No. 117, pp. 286-294

Kinoshita T., Hoshino K and Takagi K, (April 1972), Rock Breaking with Continuous High
Speed Water Jet Stream, Proceedings of 1st Int’l Symposium on Jet Cutting Technology,
Coventry, U.K.

Labus T.J., (1989), Fluid Mechanics of Jets, Fluid Jet Technology Fundamentals and
Applications, A Short Course, August 1989, Toronto Canada

Leach S.J. and Walker G..L, (1966), Some Aspects of Rock Cutting by High Speed Water
Jets, Philosphical Trans. R. Soc., Vol. 260, Series A

Lohn P.D. and Brent D.A., (April 1976), Improved Mineral Excavation Nozzle Design
Study, TRW Systems and Energy, Redondo Beach, CA, Prepared for the U.S. Bureau of
Mines

136
Mazurkiewicz M, (1983), An Analysis of One Possibility for Pulsating a High Pressure
Water Jet, Proceedings of the Second U.S. Water Jet Conference, May 24-26, 1983, Rolla
Missouri, pp17-21

Menon, S. E. (2005), Piping Calculations Manual, McGraw-Hill

Merlin Diamonds Ltd., (October 2012), Hydraulic Borehole Mining Selected Kimberlite
Pipes, Merlin Diamond Mine, Northern Territory, Australia, Prepared by Jet Mining Ltd.
and Kinley Exploration LLC

Momber, A.W., 2001, Energy Transfer during the Mixing of Air and Solid Particles into a
High-Speed Waterjet: An Impact Force Study, Exper. Thermal & Fluid Science, Vo. 25,
pp. 31-41

Nebeker E. B., (1987), Percussive Jets – State-of-the-Art, Proceedings of the Fourth U.S.
Water Jet Conference, Aug. 26-28, 1987, University of California, Berkeley, pp. 32-45

Nicklin, D.J., (1963), The Air Lift Pump: Theory and Optimization, Trans. Inst. Chem.
Eng., Vol. 41, pp. 29-39

Nikonov G.P. and Shavlovskii S.S., (1961), Gornye Mashiny i. Avtomatika, Nauchno-
Tech. Sh. 1 (8), 5

Nikonov G.P. and Goldin T.A., (1972), Coal and Rock Penetration by Fine Continuous
High Pressure Water Jets, Proceedings 1st International Symposium of Jet Cutting
Technology, Paper E2, Organized and Sponsored by BHRA Fluid Engineering, Coventry
England

Patterson, M. S. and Wong, T.F, (2005), Experimental Rock Deformation – The Brittle
Field, 2nd Ed., Springer-Verlag, Berlin and New York

Popper G., Godesky D. and Giambra, J., (1985), Phosphate Resource Potential for
Borehole Mining in the Southeastern Coastal Plain, Bureau of Mines Information Circular

Pougatch, K. and Salcudean, M., (2008), Numerical Modeling of Deep Sea Airlift, Ocean
Engineering, Vol. 35, pp. 1173-1182

137
Powell J.H. and Simpson S.P., (July 1969), Theoretical Study of the Mechanical Effects of
Water Jets Impinging on a Semi-Infinite Elastic Solid, International Journal of Rock
Mechanics and Mining Science, Vol. 6, pp. 353-364

Rehbinder G., (Nov. 1977), Slot Cutting in Rock with a High Speed Water Jet, International
Journal of Rock Mechanics and Mining Science, Vol. 14, pp. 229-234

Rehbinder G., (April 1978), Erosion Resistance of Rock, 4th International Symposium on
Jet Cutting Technology, Canterbury, U.K. , Vol. E1, pp. E1-1-E1-10

Rehbinder G., (1979), A Theory About Cutting Rock with a Water Jet, International Journal
of Rock Mechanics and Mining Science, Vol. 12, pp. 247-257

Reinemann, D.J., Parlange, J.Y. and Timmons, M.B, (1986), Theory of Small-Diameter
Airlift Flows, Int. J. of Multiphase Flow, Vol. 16, No. 1, pp. 113-122, 1990

Ripkin, J.F. and Wetzel, J.M., 1972, A Study of Fragmentation of Rock by Impingement
with Water and Solid Impactors, Final Report on U.S. Bureau of Mines, Contract HO
210021

Rouse H., Asce M, Howe J.H. and Metzler D.E., (1952), Experimental Investigations of
Fire Monitors and Nozzles, Transactions of the American Society of Civil Engineers, Vol.
117, pp. 1147-1188

Ruff, A.W. and Wiederhorn, S.M., 1979, Erosion by Solid Particle Impact, Office of Naval
Research Report NBSIR 78-1575

Savanick G., (1987), Borehole (Slurry) Mining of Coal, Uraniferous Sandstone, Oil Sands
and Phosphate Ore, Report of Investigations 9101, United States Bureau of Mines

Shavlovskii S, (1972), Hydrodynamics of High Pressure Fine Continuous Jets, Paper A6,
1st International Symposium on Jet Cutting Technology, Coventry U.K., pp. A6-81 to A6-
92.

Sheldon, G.L. and Finnie, I., 1966, The Mechanism of Material Removal in the Erosive
Cutting of Brittle Materials, J. of Engineering for Industry, pp. 393-399

Shen, B., Stephansson, O. and Rinne, M., (2014), Modeling Rock Fracturing Processes: A
Fracture Mechanics Approach Using FRACOD, Springer

138
Shima A., Takayama K., Tomita Y. and Ohsawa N., (1983), Mechanism of Impact Pressure
Generation from Spark-Generated Bubble Collapse Near a Wall, AIAA Journal, Vol. 21,
pp. 55-59

Stenning, A.H., Martin, C.B., (1968), An Analytical and Experimental Study of Air Lift
Pump Performance, J. Eng. Power Trans., ASME 90, pp. 106-110

Summers D.A., (June 1972), Water Jet Cutting Related to Jet and Rock Properties, 14th
Rock Mechanics Symposium, ASCE, pp. 569-588

Summers D.A., (1995), Waterjetting Technology, Chapman & Hall Publishers

Summers D.A., 2016, Bit Tooth Energy Blog

Tailtel, Y., Barnea, D. and Duckler, A.E., (1980), Modeling Flow Pattern Transitions for
Steady Upward Gas-Liquid Flow in Vertical Tubes, AIChE, Vol. 26, pp. 345-354

Tecen O., (Sept. 1982), High Pressure Water Jet Assisted Drag Tool Cutting of Rock
Materials, PhD. Thesis, University of Newcastle upon Tyne

Thiruvengadam A., (Dec. 1965), The Concept of Erosion Strength, Office of Naval
Research, Technical Report No. 233-9

Tiryaki B., (June 2006), Evaluation of the Indirect Measures of Rock Brittleness and
Fracture Toughness in Rock Cutting, June 2006, The Journal of the South African Institute
of Mining and Metallurgy, Vol. 106, pp. 407-423

Tursi T.P. and Deleece R.J., (1975), Development of Very High Pressure Waterjets for
Cleaning Naval Boiler Tubes, Naval Ship Engineering Center, Philadelphia Division,
Philadelphia PA

Vijay, M.M., (2001), Fluid Mechanics of Jets, Waterjet Technology Association

Wallace J.J., Price G. C. and Ackerman M.J., (1961), Hydraulic Coal Mining Research:
Experiments and Preliminary Tests, Report of Investigations 5815, Bureau of Mines,
Department of the Interior

Weber, M. and Dedegil, M.Y., (1976), Transport of Solids According to the Airlift
Principle, Proc. 4th International Conference on the Hydraulic Transport of Solids in Pipes,
Alberta, Canada, Papers H1-1-23 and X93-94

139
Witherspoon, P.A., Wang, J.S.Y., Iwai, K. and Gale, J.E., (1980), Validity of Cubic Law
for Fluid-Flow in a Deformable Rock Fracture, Water Resources Research, V16, pp. 1016-
1024

Yanaida K, (1974), Flow Characteristics of Water Jets, Proceedings 2nd International


Symposium on Jet Cutting Technology, Paper A2, Organized and Sponsored by BHRA
Fluid Engineering, Cranfield, Bedford, England.

Yoshinaga, T. and Sato, Y., (1996), Performance of an Airlift Pump for Conveying Course
Particles, Int. J. Multiphase Flow, Vol. 22, pp. 223-238

Yuan S.C. and Harrison J.P.,(2006), A Review of the State of the Art in Modeling
Progressive Mechanical Breakdown and Associated Fluid Flow in Intact Heterogeneous
Rocks, Int’l. Journal of Rock Mechanics, Vol. 43, pp. 1001-22

140
APPLICABILITY OF HYDRAULIC BOREHOLE MINING (‘HBHM’) TO
DIAMONDIFEROUS DEPOSITS
By
Daniel Morgan Boone Beck

Permission to make digital or hard copies of all or part of this work for personal and
classroom use is granted without fee provided that copies are not made or distributed for
profit or commercial advantage and that copies bear this notice and the full citation on the
last page. To copy otherwise, to republish, to post on servers or to redistribute to lists,
requires prior specific permission and may require a fee.

141

S-ar putea să vă placă și